Hertfordshire Chalk PDF
Hertfordshire Chalk PDF
Hertfordshire Chalk PDF
2010
2
I, Ciara M. Fitzpatrick, confirm that the work presented in this thesis is my own. Where information
has been derived from other sources, I confirm that this has been indicated in the thesis.
3
Foreword
This Engineering Doctorate (EngD) was funded by an Engineering and Physical Sciences Research
Council (EPSRC) Studentship in association with Veolia Water Three Valleys Ltd1 (VWTV) and Thames
Water Utilities Ltd (TWUL).
As part of the research project two EngD studentships were awarded: (1) to Ciara Fitzpatrick, orig-
inator of this document, and (2) to Simon Cook. The research project was intended to be collaborative
between the water utilities and University College London, with neither individual student affiliated
preferentially to either company.
The work documented within this thesis constitutes independent original research, however the
research described is complementary to that by Simon Cook in:
The views, opinions and conclusions expressed in the thesis are those of the researcher and do not
necessarily reflect the views of Veolia Water Three Valleys Ltd or Thames Water Utilities Ltd (TWUL).
Abstract
Bromate contamination over an area of more than 40 km2 in the Hertfordshire Chalk aquifer was first
detected in 2000 and is the largest case of point-source groundwater contamination in the UK. Bromate
is a possible human carcinogen, and a regulatory limit for drinking water of 10 µg l−1 had been imple-
mented in the U.K. since 2003. Background concentrations of bromate in groundwater are believed to
be effectively zero. In the affected area, bromate at concentrations of several 100 µg l−1 have forced
the closure of a large public water supply source and restricted the use of seven other public supply
boreholes up to 20 km from the contamination source.
The source has been identified as a former industrial site which operated between 1955 and 1983.
Residual contamination at the site provides a continuing source of bromate to groundwater. A range
of conceptual scenarios for bromate mobilisation and release to groundwater have been developed and
quantified based on interpretation of the available data, and constrained by interpolation of the observed
concentrations.
Analysis and interpretation of all available monitoring and investigation data throughout the catch-
ment has revealed the influence of recharge, water level, and groundwater abstractions on bromate con-
centrations. These relationships, integrated with observations of the geology and hydrogeology of the
area, support a conceptualisation of transport of bromate by dominantly double-porosity processes within
the Vale of St. Albans area, which maintains a highly attenuated, stable contaminant distribution west of
Hatfield. An extensive karst network related to the position of the Palaeogene overlap of the Chalk influ-
ences bromate transport to the east of Hatfield, dispersing bromate rapidly over large distances toward the
Lea Valley. The revised conceptual understanding has enabled the development of a new interpretation
of bromate transport within the catchment between 2000 and 2008.
A new analytical network modelling approach has been developed to predict the long-term, large-
scale transport of bromate. The model simulates Fickian double-porosity diffusive exchange along in-
terconnecting flow-lines, linked to rapid karst flow. The model is parameterised on the basis of single
borehole dilution testing, catchment-scale natural gradient tracer testing, and literature derived values.
The network model, combined with quantified bromate source terms, simulates bromate and bromide
concentrations of the order of magnitude of those observed at locations within the Vale of St. Albans, and
predicts bromate concentrations to remain above regulatory limits for around 200 years. This highlights
the importance of double-porosity diffusion for the long-term evolution of contaminants at catchment-
scale in the Chalk aquifer.
5
Acknowledgements
I am sincerely grateful to my supervisors, Dr Willy Burgess and Prof. John Barker for providing inspi-
ration, support and guidance over the past four years, and above all, for believing in my capabilities. I
am indebted to Willy for his constant enthusiasm, and ability to renew mine.
This Engineering Doctorate was funded by an EngD studentship at UCL and I am grateful to EP-
SRC, Veolia Water Three Valleys Ltd, and Thames Water Utilities Ltd for sponsoring the project. In
particular, I would like to thank my supervisors Rob Sage, Lucy Lytton, and Philip Bishop for their
insights and information.
I owe my gratitude to Jon Newton at the Environment Agency, firstly for introducing me to the
Hertfordshire bromate problem, and subsequently for sharing his understanding, providing monitoring
data and a wealth of other relevant information.
The single borehole dilution tests were undertaken with assistance from a number of people: many
thanks to the Environment Agency for permission and funding, Adrian Sheriff for access to the Nashe’s
Farm borehole, Simon Cook and Gemma Russell for help with the fieldwork, Louise Maurice for guid-
ance on the method and use of the equipment, and Jessica Randle for associated geophysical testing of
the boreholes.
I am grateful to Simon Cook, for his contributions to understanding and progressing research into
the bromate contamination. Thanks also to Rakia Meister, Simon Quinn, Mohammed Abdul Hoque,
Qiong Li, Mike Davis, Bethan Hallett and Gemma Russell for putting up with sharing an office with me,
and for some welcome conversation.
‘Go raibh maith agat’ to my fiancé, Neal O’Grady, for his love, understanding and patience. Finally,
I thank my family; without their love and support over the last 29 years I would not be in a position to
be submitting this thesis. I dedicate this thesis to my grandparents, Mona and Chris Fitzpatrick.
6
Abbreviations
BH Borehole
EA Environment Agency
TVW Three Valleys Water Limited (renamed Veolia Water Three Valleys Limited in 2009)
Contents 7
Contents
1 Introduction 22
1.1 Background . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22
1.2 Bromate transport in the Hertfordshire Chalk aquifer . . . . . . . . . . . . . . . . . . . 24
1.3 The bromate source . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 25
1.4 Research aims and objectives . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 25
1.5 Approach . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 26
1.6 Structure of thesis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 27
1.7 Environmental Hydrochemistry of Bromate and Bromide . . . . . . . . . . . . . . . . . 27
1.7.1 Occurrence of Bromate and Bromide in surface and groundwaters . . . . . . . . 27
1.7.2 Environmental Behaviour . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 29
3 A conceptual model for flow and transport of bromate in the Hertfordshire Chalk 62
3.1 Chapter Objective . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 62
3.2 Geology and Hydrogeology of Hertfordshire . . . . . . . . . . . . . . . . . . . . . . . . 62
3.2.1 Topography . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 62
3.2.2 Hydrology . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 62
3.2.3 Geology . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 64
3.2.4 Hydrogeology . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 69
3.2.5 Karstic Features . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 72
3.2.6 Karst Flow . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 73
3.2.7 Groundwater–surface water interactions . . . . . . . . . . . . . . . . . . . . . . 77
3.2.8 Chalk–Drift Groundwater interactions . . . . . . . . . . . . . . . . . . . . . . . 78
3.3 Piezometry . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 78
3.3.1 Groundwater flow in the Sandridge-St Leonard’s Court Area . . . . . . . . . . . 83
3.3.2 Abstractions and Discharges . . . . . . . . . . . . . . . . . . . . . . . . . . . . 84
3.4 Regional hydrochemisty . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 84
3.5 Scavenge Pumping at Hatfield Pumping Station . . . . . . . . . . . . . . . . . . . . . . 85
3.5.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 85
3.5.2 Data sources . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 86
3.5.3 Data handling . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 86
3.5.4 Abstraction rates at Hatfield . . . . . . . . . . . . . . . . . . . . . . . . . . . . 86
3.5.5 Bromate and Bromide time series trends . . . . . . . . . . . . . . . . . . . . . . 86
3.5.6 Bromate and Bromide: Relationship to Hatfield abstraction rates . . . . . . . . . 100
Contents 9
6.2.5 Catchment-scale distributed flow modelling using MODFLOW and MT3D . . . 240
6.2.6 Weaknesses of MODFLOW and MT3D . . . . . . . . . . . . . . . . . . . . . . 242
6.3 Development of a Multiple Analytical Pathways Approach . . . . . . . . . . . . . . . . 245
6.3.1 DP1-D . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 245
6.3.2 Multiple Analytical Pathways (MAP) model . . . . . . . . . . . . . . . . . . . 245
6.3.3 GoldSim Contaminant Transport Model . . . . . . . . . . . . . . . . . . . . . . 248
6.3.4 Comparison of DP1D, MAP and GoldSim . . . . . . . . . . . . . . . . . . . . . 248
6.4 Analytical Network Model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 249
6.4.1 Mathematical Basis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 249
6.4.2 Node Input - the Source Function . . . . . . . . . . . . . . . . . . . . . . . . . 249
6.4.3 Node and Branch description . . . . . . . . . . . . . . . . . . . . . . . . . . . . 249
6.4.4 Node and Branch Output . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 250
6.5 A Network Model for Hertfordshire . . . . . . . . . . . . . . . . . . . . . . . . . . . . 250
6.5.1 Selection of Nodes and Branches . . . . . . . . . . . . . . . . . . . . . . . . . 250
6.5.2 Parameters for ‘double-porosity’ branches . . . . . . . . . . . . . . . . . . . . . 250
6.5.3 Parameters for karst branches . . . . . . . . . . . . . . . . . . . . . . . . . . . 253
6.6 Network Model for Hertfordshire - Results of initial simulations . . . . . . . . . . . . . 254
6.7 Discussion and conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 256
7 Conclusions 268
7.1 Fulfillment of research aims and objectives . . . . . . . . . . . . . . . . . . . . . . . . 268
7.1.1 Evolution of bromate contamination . . . . . . . . . . . . . . . . . . . . . . . . 268
7.1.2 The source . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 269
7.1.3 Catchment-scale modelling of bromate transport . . . . . . . . . . . . . . . . . 270
7.2 Recommendations for further work . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 271
Bibliography 272
Appendices 281
List of Tables
1.1 Bromide in UK groundwaters, summarised from Edmunds et al. (1989). r2 is the linear
correlation coefficient squared for a regression of Br vs. Cl . . . . . . . . . . . . . . . . 28
2.1 Molecular diffusion coefficients in fissured and unfissured chalk. After Hill (1984).
These values represent the mass flux through the saturated matrix per unit concentra-
tion gradient in the water. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 54
2.2 Characteristic times for infinite slab geometry, with slabs of thickness 2b separated by
fractures of aperture a. For this model, the ratio of volume to area for a matrix block (`)
is represented by b. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 60
4.1 Typical analytical suite for water samples May 2000 to December 2008 . . . . . . . . . 120
4.2 Analytical methodology and detection limits for bromate analyses. . . . . . . . . . . . 120
4.3 Regression statistics for the response of bromide concentration to bromate concentration
and chloride concentration. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 162
5.1 Chronology and scope of site investigations and monitoring at the source site . . . . . . 171
5.2 Geological strata encountered at the source site. Based on Komex (2000) and Atkins
(2002) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 174
List of Tables 13
5.3 Summary of mass estimates. Estimates for total mass in the unsaturated and saturated
zones refer to minimum, mean and maximum thicknesses defined in Figure 5.31. . . . . 207
5.4 Main areas of uncertainty in the history of bromide and bromate release to groundwater
beneath the source site. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 211
5.5 Mass predicted by source history scenarios A and B compared to observed mass con-
straints. Condition 4 is based on an estimate by Buckle (2002) of the mass removed at
Hatfield and Essendon between 1981 and 2000. . . . . . . . . . . . . . . . . . . . . . . 222
6.1 Parameter combinations for ‘best-case’ (lowest peak bromate concentrations) and
‘worst-case’ (highest peak bromate concentrations) scenarios. . . . . . . . . . . . . . . 253
6.2 Parameters derived from fitting the DP-1D model (Barker, 2005) to tracer breakthrough
curves from Water End injection (Cook, 2010). Characteristic times are in hours. . . . . 254
List of Figures 14
List of Figures
1.1 Extent of the bromate contamination in Hertfordshire. The regulatory limit for bromate
in drinking water is 10 µg l−1 . Background concentrations are effectively zero. . . . . . 23
1.2 Plot of Br vs. Cl concentrations in groundwaters from the Chalk of the Colne and Lee
River catchments. Points plotting in the ‘contamination’ box are from the Hatfield area.
From Shand et al. (2003). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 28
2.1 Relationship between the traditional and revised stratigraphy of the Chalk. After
Woods and Aldiss (2004) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 31
2.2 (a) An idealised double-porosity aquifer; (b) an idealised double-permeability aquifer.
From Price et al. (1993) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 39
2.3 Relationsip between fissure spacing, aperture, porosity and hydraulic conductivity for
a fissure system containing three plane, parallel, mutually perpendicular smooth-walled
fissures filled with pure water at 10 degC and porosity relationship (Price et al., 1993). . 44
2.4 Ranges of fracture spacings and fracture apertures for the Chalk. Results as cited in
Bloomfield (1996) and Watson (2004). . . . . . . . . . . . . . . . . . . . . . . . . . . . 46
2.5 Double-porosity diffusive exchange of solutes. At an early stage, diffusion from the con-
taminated fracture water into the matrix water acts to retard the transport of contaminants
down-gradient. At a later stage, contaminated porewater acts as a persistent secondary
source of contamination. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 53
2.6 Governing equations and assumptions for a double-porosity mocel with slab geometry.
After Barker (1982). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 59
3.8 Average piezometry 1998 to 2008. Contour levels are in m AOD. Arrows indicate
groundwater flow direction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 82
3.9 Abstraction rates at Hatfield PS between 31 June 2005 and 31 December 2008 . . . . . . 87
3.10 Time series of bromate and bromide concentrations at Hatfield PS, soil moisture deficit,
and monthly rainfall. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 88
3.11 Time series of bromate and bromide concentrations at Essendon PS, soil moisture deficit,
and monthly rainfall. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 89
3.12 Time series of bromate and bromide concentrations at Chadwell Spring, soil moisture
deficit, and monthly rainfall. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 90
3.13 Time series of bromate and bromide concentrations at Amwell Hill PS, soil moisture
deficit, and monthly rainfall. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 91
3.14 Time series of bromate and bromide concentrations at Amwell Marsh PS, soil moisture
deficit, and monthly rainfall. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 92
3.15 Time series of bromate and bromide concentrations at Rye Common PS, soil moisture
deficit, and monthly rainfall. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 93
3.16 Time series of bromate and bromide concentrations at Middlefield Road PS, soil mois-
ture deficit, and monthly rainfall. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 94
3.17 Time series of bromate and bromide concentrations at Hoddesdon PS, soil moisture
deficit, and monthly rainfall. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 95
3.18 Time series of bromate and bromide concentrations at Broxbourne PS, soil moisture
deficit, and monthly rainfall. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 96
3.19 Time series of bromate and bromide concentrations at Turnford PS, soil moisture deficit,
and monthly rainfall. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 97
3.20 Assessment of residuals for each ’best-fit’ regression for the response of bromate con-
centration to Hatfield abstraction rate. (1) . . . . . . . . . . . . . . . . . . . . . . . . . 103
3.21 Assessment of residuals for each ’best-fit’ regression for the response of bromate con-
centration to Hatfield abstraction rate. (2) . . . . . . . . . . . . . . . . . . . . . . . . . 104
3.23 Comparison of statistical response times for bromate concentration response to hatfield
abstraction and tracer travel times from Water End. Based on Cook (2010) . . . . . . . . 109
3.24 Methodology for determination of specific discharge (darcy velocity) from the results of
the Single Borehole Dilution Tests. Based on Ward et al. (1998) . . . . . . . . . . . . . 111
3.25 Specific discharge (darcy velocity) for each 0.5 m depth section at Nashes Farm. Es-
Ct −Cb
timated using the methodology in Figure 3.24. Plots of ln C 0 −Cb
are included in Ap-
pendix D. The value at each section is estimated based on Based on Single Borehole
Dilution test carried out at Nashes Farm 29 January 2008. . . . . . . . . . . . . . . . . 112
List of Figures 16
3.26 Specific discharge (darcy velocity) for each 0.5 m depth section at Comet Way BH.
Ct −Cb
Estimated using the methodology in Figure 3.24. Plots of ln C0 −Cb
are included in
Appendix D. The value at each section is estimated based on Based on Single Borehole
Dilution test carried out at Comet Way BH 4 February 2008. . . . . . . . . . . . . . . . 113
3.27 Specific discharge (darcy velocity) for each 0.5 m depth section at Harefield House BH.
Ct −Cb
Estimated using the methodology in Figure 3.24. Plots of ln C0 −Cb
are included in
Appendix D. The value at each section is estimated based on Based on Single Borehole
Dilution test carried out at Harefield House BH on 22 January 2008. . . . . . . . . . . . 114
3.28 Conceptual model for groundwater flow in the bromate affected area of Hertfordshire.
Position of conduits are based on the conceptual model developed by Cook (2010). Flow
rates and attenuation characteristics are inferred from the results of the single borehole
dilution testing presented in Section 3.6 and tracer tests undertaken by Cook (2010). . . . 115
4.23 Time series of bromate and bromide concentrations at selected locations in the Hatfield
Quarry area, soil moisture deficit, and monthly rainfall. . . . . . . . . . . . . . . . . . . 146
4.24 Time series of bromate and bromide concentrations at selected locations in the Hatfield
Quarry area, soil moisture deficit, and monthly rainfall. . . . . . . . . . . . . . . . . . . 147
4.25 Time series of bromate and bromide concentrations at selected locations in the Hatfield
Quarry area, soil moisture deficit, and monthly rainfall. . . . . . . . . . . . . . . . . . . 148
4.26 Time series of bromate and bromide concentrations at selected locations in the Hatfield
area, soil moisture deficit, and monthly rainfall. . . . . . . . . . . . . . . . . . . . . . . 151
4.27 Time series of bromate and bromide concentrations at selected locations in the Hatfield
area, soil moisture deficit, and monthly rainfall. . . . . . . . . . . . . . . . . . . . . . . 152
4.28 Time series of bromate and bromide concentrations at selected locations in the Hatfield
area, soil moisture deficit, and monthly rainfall. . . . . . . . . . . . . . . . . . . . . . . 153
4.29 Time series of bromate and bromide concentrations at selected locations in the Hatfield
area, soil moisture deficit, and monthly rainfall. . . . . . . . . . . . . . . . . . . . . . . 154
4.30 Time series of bromate and bromide concentrations at selected locations in the Hatfield
area, soil moisture deficit, and monthly rainfall. . . . . . . . . . . . . . . . . . . . . . . 155
4.31 Time series of bromate and bromide concentrations at selected locations in the Lea Val-
ley, soil moisture deficit, and monthly rainfall. . . . . . . . . . . . . . . . . . . . . . . . 157
4.32 Annual average Bromide concentrations in groundwater 2000 to 2008. . . . . . . . . . . 159
4.33 Bromide concentrations at locations where bromate concentrations are less than MDL. . 160
4.34 Bromate/Bromide ratio variation with bromate concentration for groundwater and sur-
face water samples . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 161
4.35 Spatial distribution of mean annual bromate/bromide ratio 2000 to 2008. . . . . . . . . . 163
4.36 Percentage of samples of bromate concentrations for which there are accompanying wa-
ter level measurements for each location . . . . . . . . . . . . . . . . . . . . . . . . . . 164
4.37 Regression relationship for the response of bromate concentration to water level. Per-
centages refer to the amount of variation explained by the regression (R2 value) . . . . . 165
5.1 Location of the source site in Sandridge, Hertfordshire. Formerly the Steetly chemical
works, now the St Leonard’s Court residential development. . . . . . . . . . . . . . . . 168
5.2 Location of former process areas of the Steetly Chemical Works. Based on Atkins (2002)
interpretation of historical plans, aerial photographs, and the interview with a former
employee of the works. Aerial photograph taken in 1971. . . . . . . . . . . . . . . . . . 170
5.3 Borehole locations from investigations 1983-1985 (STATS, 1983a,b,c, 1984; Chemfix,
1985c) and 2000-2001 (Komex, 2000; Atkins, 2002). For locations from 1983-1985,
numbers in square brackets indicate date of drilling. . . . . . . . . . . . . . . . . . . . . 172
5.4 Trial hole locations from investigations in 1985 (Chemfix, 1985c) . . . . . . . . . . . . 173
5.5 Piezometry at the St Leonard’s Court site November 2001. From Atkins (2002) . . . . . 175
5.6 Cross-section parallel to groundwater flow direction. From Atkins (2002) . . . . . . . . 176
List of Figures 18
5.26 Groundwater monitoring locations in the vicinity of the source site that have been sam-
pled for bromide concentrations between 1983 and 1987 and between 2000 and 2008. . . 199
5.27 Groundwater bromide concentrations at monitoring locations in the vicinity of the source
site 1983 to 2008. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 200
5.28 Relationships between leachate concentration (mg l−1 ) and soil concentration
(mg kg−1 ) for samples from the 2001 investigation (Atkins, 2002). . . . . . . . . . . . . 201
5.29 Bromate soil concentration contours for 1.0 m thick grid slices based on investigation
data from 2000 and 2001 (Komex, 2000; Atkins, 2002). Estimates for total mass in
the unsaturated and saturated zones refer to minimum, mean and maximum thicknesses
defined in Figure 5.31. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 204
5.30 Bromide soil concentration contours for 1.0 m thick grid slices based on investigation
data from 2000 and 2001 (Komex, 2000; Atkins, 2002). Estimates for total mass in
the unsaturated and saturated zones refer to minimum, mean and maximum thicknesses
defined in Figure 5.31. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 205
5.31 Minimum and maximum saturated and unsaturated zone thicknesses. . . . . . . . . . . . 206
5.32 Estimates of bromate groundwater flux from the ‘source zone’ using equation 5.2 and
the area under a concentration profile taken across a flux plane through the source zone.
R x=B
The area under a graph represents the integral x=A C dx. The flux plane is shown in
Figure 5.23. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 209
5.33 The combined ‘source zone’ (centre figure) based on the locations of high concentrations
of bromate (left hand figure) and bromide (right hand figure) in groundwater . . . . . . . 212
5.34 Conceptual Model for bromate and bromide release from the source zone. . . . . . . . . 213
5.35 Derivation of equations for mass of bromide/bromate in the unsaturated zone and the
rate of input of bromide/bromate from the unsaturated zone to the saturated zone. . . . . 218
5.36 Equations for bromide mass, fit to observed values from 1985 and 2001. Parameters are
defined in Figure 5.11.1 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 220
5.37 Bromide and bromate concentrations for Scenario A and Scenario B from 1984 into the
future. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 221
5.38 Bromide source history for Scenarios A and B. . . . . . . . . . . . . . . . . . . . . . . 223
5.39 Bromide source history for Scenarios A and B. . . . . . . . . . . . . . . . . . . . . . . 224
5.40 Bromide and bromate concentrations for Scenario A and Scenario B between 1955 and
1984. After 1984 concentrations proceed as in Figure 5.37. . . . . . . . . . . . . . . . . 225
5.43 Comparison of simulated bromide concentrations for source history Scenario A and ob-
served concentrations at three monitoring locations. . . . . . . . . . . . . . . . . . . . . 232
5.44 Comparison of simulated bromate concentrations for source history Scenario A and ob-
served concentrations at three monitoring locations. . . . . . . . . . . . . . . . . . . . . 233
5.45 Comparison of simulated bromide concentrations for source history Scenario B and ob-
served concentrations at three monitoring locations. . . . . . . . . . . . . . . . . . . . . 234
5.46 Comparison of simulated bromate concentrations for source history Scenario B and ob-
served concentrations at three monitoring locations. . . . . . . . . . . . . . . . . . . . . 235
5.47 Comparison of simulated bromate concentrations for source history Scenario C and ob-
served concentrations at three monitoring locations. . . . . . . . . . . . . . . . . . . . . 236
5.48 Concurrent matrix and fissure concentrations are required to determine at which point
along the concentration-time graph a particular fissure concentration represents. . . . . . 237
6.1 Comparison of the superseded versions of the the source terms used by Cook (2010) to
the current versions in this thesis. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 243
6.2 Conceptual and mathematical basis for the Multiple Analytical Pathways of Barker (2001).246
6.3 Conceptual and mathematical basis for the Multiple Analytical Pathways of Barker (2001).247
6.4 Nodes and branches represented in the Network Model for Hertfordshire. Note that
branches are shown schematically as straight-line connectors and are not intended to
indicate the precise geographical route. . . . . . . . . . . . . . . . . . . . . . . . . . . . 251
6.5 Simulated bromate concentrations at Harefield House using source terms for Scenario
A, B and C (Section 5.11), and a constant concentration source term of 5000 µg l−1 . . . 257
6.6 Simulated bromate concentrations at Hatfield Quarry using source terms for Scenario A,
B and C (Section 5.11), and a constant concentration source term of 5000 µg l−1 . . . . . 258
6.7 Simulated bromate concentrations at Comet Way using source terms for Scenario A, B
and C (Section 5.11), and a constant concentration source term of 5000 µg l−1 . . . . . . 259
6.8 Simulated bromate concentrations at Arkley Hole Spring node, and at the end of con-
tributing branches, using source terms for Scenario A, B and C (Section 5.11), and a
constant concentration source term of 5000 µg l−1 . . . . . . . . . . . . . . . . . . . . . 260
6.9 Simulated bromate concentrations at Lynchmill Spring node, and at the end of contribut-
ing branches, using source terms for Scenario A, B and C (Section 5.11), and a constant
concentration source term of 5000 µg l−1 . . . . . . . . . . . . . . . . . . . . . . . . . . 261
6.10 Simulated bromide concentrations at Harefield House using source terms for Scenario A
and B. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 262
6.11 Simulated bromide concentrations at Hatfield Quarry using source terms for Scenario A
and B. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 263
6.12 Simulated bromide concentrations at Comet Way using source terms for Scenario A and B.264
6.13 Simulated bromate concentrations for Scenario C at Harefield House, Hatfield Quarry,
and Comet Way using ‘best-case’, ‘typical-case’ and ‘worst-case’ parameters. . . . . . . 265
List of Figures 21
6.14 Concurrent matrix and fissure concentrations are required to determine at which point
along the concentration-time graph a particular fissure concentration represents. . . . . . 266
22
Chapter 1
Introduction
1.1 Background
Changes to the Water Supply (Water Quality) Regulations (2000), effective from 2003, introduced bro-
mate (BrO3– ) as a new sampling parameter with a regulatory limit of 10 µg l−1 in drinking water. Bro-
mate is a possible human carcinogen based on extrapolation from rodent studies. Background concen-
trations of bromate in groundwater are believed to be effectively zero. In May 2000, during the course of
preliminary sampling in advance of the regulations, Three Valleys Water Ltd1 (TVW), detected bromate
concentrations of 135-140 µg l−1 , well in excess of this standard, at the Hatfield Bishop’s Rise Pumping
Station. The Environment Agency (EA) and the Drinking Water Inspectorate (DWI) were informed and
the Hatfield source was removed from public supply.
In June 2000 a joint water quality monitoring programme was initiated, involving the Environment
Agency, the local authorities, and the water companies to identify the source and extent of the bromate
contamination. The source has been identified as a former industrial site in Sandridge (Figure 1.1) which
operated between 1955 and 1983. The site is now the St Leonard’s Court residential development. This
site has been determined as ‘Contaminated Land’ and designated a ‘Special Site’ as defined under Part
IIA of the Environmental Protection Act 1990. The bromate contamination was found to extend up to
20 km to the east of the source site affecting an area of more than 40 km2 of the Hertfordshire Chalk
aquifer (Figure 1.1), and restricting the use of a further seven public water supply boreholes in the Lea
Valley. Bromide (Br – ) concentrations were also found to be elevated above background concentrations
across the catchment. The Environment Agency, TVW and TWUL have continued to monitor water
quality and water levels at a number of locations throughout the area affected by the bromate plume. A
chronology of key events is included in Appendix A.
The Chalk is one of the most important aquifers in the UK; according to the UK Groundwater
Forum2 it provides over half of all groundwater for public supply in the UK and, in the south-east of
England, up to 70 % of all water in public supply. The area of Chalk aquifer affected by the bromate
contamination has a licensed abstraction of approximately 200 Ml d−1 . Surface waters of the River Lea,
and tributaries of the River Colne are also affected. The bromate contamination is therefore a significant
1 renamed in 2009 as Veolia Water Three Valleys Limited
2 www.groundwateruk.org/
1.1. Background 23
Figure 1.1: Extent of the bromate contamination in Hertfordshire. The regulatory limit for bromate in
drinking water is 10 µg l−1 . Background concentrations are effectively zero.
1.2. Bromate transport in the Hertfordshire Chalk aquifer 24
threat to the long-term quality of a number of strategic public water supply sources, and also many private
supply sources. The financial cost to the water companies, which include costs incurred for additional
monitoring of sources, treatment of bromate and bromide contaminated water, operation of the interim
scavange pumping at Hatfield PWS, and drilling and investigatory work for replacement supply wells,
have been estimated to be of the order of in the region of £50,000,000 for TVW and TWUL (R. Sage &
P. Bishop, pers. comm.) for the period 2000 to 2006. A Business Case for the reserach is included in
Appendix B.
Such a system would allow rapid transport of bromate, with low-attenuation of concentrations, and is
likely to be the cause of the wide distribution of bromate across the east of the catchment. Previous
modelling exercises (e.g. Buckle, 2003; Atkins, 2005) have been unable to reproduce the migration of
bromate to the Lea Valley due to deficiencies in the representation of the karst system.
Therefore, in order to represent the distribution and temporal evolution of bromate contamina-
tion in the Hertfordshire Chalk, predictive models of bromate transport must integrate the effects of
double-porosity diffusive exchange between fissures and the chalk matrix with the effects of rapid, low-
attenuation transport within karstic conduits. The complex nature of these processes, and quantifying
their relative importance, presents particular difficulties for prediction of contaminant transport.
• To use the available information and monitoring data to describe the spatial distribution and tem-
poral evolution of bromate across the catchment, and to interpret this in association with the con-
ceptual model of the flow and transport system.
1.5. Approach 26
The source
• To describe and quantify the distribution of bromate at the source site through collation and de-
scription of site investigation and monitoring data;
• To develop alternative conceptual scenarios for bromate release to groundwater and quantify these
as ‘source terms’;
• To use the available monitoring data to constrain the potential source terms.
• To use this model to produce predictions for the likely bromate concentrations at key output loca-
tions over the long-term.
1.5 Approach
The objectives were approached by developing a conceptual model of groundwater flow within the
Hertfordshire Chalk, which was informed by a new interpretation of the Hertfordshire karst system
by Cook (2010) following catchment-scale tracer tests. All available monitoring and investigation data
were analysed and interpreted to describe the evolution of the bromate contamination within the aquifer.
The processes controlling bromate transport were investigated by examining relationships between bro-
mate concentrations and variables including bromide and chloride concentrations, piezometry, catchment
recharge, and groundwater abstraction rates.
The bromate pollution in Hertfordshire is of a similar scale to the extensive chloride pollution of the
Chalk of the Tilmanstone valley in Kent (Watson, 2004) from coalfield brines. Watson (2004) developed
an investigatory methodology - the ‘Tilmanstone Methodology’ - for characterising and parameterising
catchment-scale double-porosity flow and transport in the Chalk. The investigatory methodology com-
prised geophysical testing, single borehole dilution tracer testing, borehole-to-borehole natural gradient
tracer testing, and sampling of chalk porewater and chalk fissure water to produce vertical porewater
and fracture water profiles at locations along the flow line. A one-dimensional semi-analytical model
(DP1D), incorporating Fickian diffusion between matrix water and fracture water, was used to simulate
chloride migration and was able to reproduce the observed porewater and fracture water profiles. For-
ward modelling indicated that the double porosity diffusion extends the duration of contamination in the
catchment by several decades.
It was therefore considered that the ‘Tilmanstone Methodology’ could be adapted for application
to the Hertfordshire bromate contamination. However, initial plans for a number of cored boreholes to
provide porewater profiles in the catchment had to be reconsidered due to the financial constraints of the
Water Companies. The scope of the field testing was reduced to a number of single borehole dilution
tests to determine groundwater velocities in the Hertfordshire Chalk.
1.6. Structure of thesis 27
Representing Fickian double-porosity diffusion was still considered of great importance, despite the
lack of data for porewater concentrations to fully validate predictions of fissure concentrations. There-
fore, an analytical network model was developed which represents Fickian double-porosity diffusive
exchange between fissure water and matrix porewater along interconnecting flow-lines, while allowing
karstic branches to be incorporated into the network. The model was parameterised by a combination of
values found within the literature, and the results of the single borehole dilution testing and catchment-
scale natural gradient tracer testing. Results were compared to groundwater monitoring data at key
locations within the catchment.
The bromate source term for the Hertfordshire network model was quantified through analysis and
interpretation of site investigation data available for the source site to constrain a range of conceptual
scenarios for bromate mobilisation and release to groundwater.
potable groundwaters in the United Kingdom range from 60 to 340 µg l−1 . Concentrations in the Chalk
are summarised in Table 1.1. Due to the similarity of the geochemical behaviour of Br and Cl, using the
Br/Cl ratio is necessary to identify significant anomalies in the natural environment (Edmunds, 1996).
Table 1.1: Bromide in UK groundwaters, summarised from Edmunds et al. (1989). r2 is the linear
correlation coefficient squared for a regression of Br vs. Cl
In a survey of baseline quality of the Chalk of the Colne and Lee River catchments, Shand et al.
(2003) identified elevated concentrations of Br – : samples from the vicinity of Hatfield had (Br)/(Cl) ra-
tios which plotted above the background trend (Figure 1.2).
Figure 1.2: Plot of Br vs. Cl concentrations in groundwaters from the Chalk of the Colne and Lee River
catchments. Points plotting in the ‘contamination’ box are from the Hatfield area. From Shand et al.
(2003).
In contrast to bromide, bromate (BrO3– ) is not reported as occurring naturally in surface waters
and is not normally present in aquifers (Butler et al., 2005). However, bromate has been detected in the
surface water environment as a result of industrial oxidation/disinfection processes (Butler et al 2005)
1.7. Environmental Hydrochemistry of Bromate and Bromide 29
with one study of 36 river samples detecting bromate at levels from 4 to 8 µg l−1 (Kruithof and Meijers,
1995). The occurrence of bromate within aquifers other than the Hertfordshire Chalk could not be found
reported within the published literature.
3 KBrO
3 solubility at 25◦ C is 75 g l−1
30
Chapter 2
2.1 Introduction
The aim of this chapter is to describe the mechanisms of groundwater flow and transport of solutes
within the Chalk, with particular reference to behaviour of the Hertfordshire Chalk aquifer, to review
typical parameters related to flow and transport in the Hertfordshire Chalk, and to review approaches to
representation and modelling of flow and transport in the Chalk.
Three subdivisions were traditionally recognised in the Chalk Group: Lower, Middle and Upper Chalk.
Over the last 20 years there has been much research into Chalk Group stratigraphy (e.g. Bristow et al.
(1997); Gale et al. (1999); Woods (2006)) and the British Geological Survey has now adopted a revised
classification that follows Rawson et al. (2001) (Figure 2.1).
2.2.2 Lithology
In general the Chalk is a very fine grained (less than 10 µm), pure (circa 98 % CaCO3 ), soft, white
limestone of high (∼40 %) interstitial porosity, containing some marl bands and flint (Hancock, 1975),
with pervasive fractures of a variety of styles. The Chalk matrix is composed of calcium carbonate
micro-fossil fragments: the major, finer fraction comprising coccoliths and coccolith debris, and the
minor, coarser fraction comprising foraminifera and other shell debris (Hancock, 1975).
The marl bands can be several centimetres thick, some of them being laterally continuous for several
hundreds of kilometres. Many were deposited on erosional surfaces and are thought to be of volcanic
origin as they contain Mg-rich smectite. Flint occurs predominantly in layers parallel to bedding, either in
tabular layers or as scattered discrete nodules. Locally it forms cross-cutting veins and vertical cylinders
with a burrow at the core. ‘Hardgrounds’ are horizons of hard, brittle chalk in which porosity is reduced
to 10-20 %. The formation of hardgrounds is thought to be a result of sea floor cementation associated
with a reduction or cessation in the supply of coccoliths.
2.2. The Chalk 31
Figure 2.1: Relationship between the traditional and revised stratigraphy of the Chalk. After Woods and
Aldiss (2004)
2.3. The Chalk as an aquifer 32
interstices of the rock matrix, but the pore spaces are so fine that this water is effectively immobile. The
mobile water (the remaining 5 % or less) is held within the fractures that transect the chalk matrix. Some
fractures have been enlarged by dissolution to become fissures or even karstic conduits. The fractures,
fissures and conduits provide the permeable pathways for flow.
Barker (1993) provided illustrative relative flow rates permitted by matrix, primary fissures (frac-
tures) and secondary fissures (fractures enlarged by solution): rates of movement of groundwater through
the Chalk’s secondary fissure system2 can often be several hundreds of metres per day, compared to rates
of the order of a few millimetres per day in the primary fissure system3 , and the order of just a millimetres
per year in the matrix4 . The matrix water is therefore considered to be essentially immobile.
varying degrees of karst, rather than being wholly karstic or wholly non-karstic (Atkinson and Smart,
1981), or that all carbonate aquifers should be considered as triple-porosity/permeability aquifers, with
matrix, fissure and channel/conduit components (Worthington, 2003).
MacDonald et al. (1998) presented various strands of evidence which point to rapid groundwater
flow being widespread throughout the Chalk of Southern England. This evidence was drawn from:
• Geomorphological features;
• Rapid flow rates implied by tracer tests, including those associated with observed karstic features
(such as swallow holes, stream sinks, springs etc) and those carried out without direct association
to such features (such as boreholes, soakaways);
The evidence suggests that rapid groundwater flow is generally more frequent close to Palaeogene
cover and may also be associated with other forms of cover and valley bottoms (Section 2.4.3).
Swallow hole The point at which a stream sinks underground. A swallow hole generally implies nearly
instantaneous water loss into an opening at the bottom of a sinkhole or karst valley, whereas a
swallet may refer to gradual water loss into the gravel along a streambed, with no depression
apparent.
Sinkhole A closed depression (circumscribed by a closed topographic contour) which drains to the
subsurface.
Doline (dissolution sinkhole) A sinkhole resulting from gradual dissolution of the bedrock. A classic
bowl-shaped contour (gently sloping depression that is wider than it is deep). Dolines can be
considered as depressions that do not necessarily have water flowing into them.
Sinking stream any stream that disappears underground, typically into a swallow hole.
Spring Any natural discharge of water from rock or soil onto the surface of the land of into a body or
surface water. Springs which occur on the dip slope of the Chalk are nearly always in the bottoms
of valleys and reflect the emergence of the water table at the surface.
2.4. Karstic behaviour of the Chalk 35
Bourne Seasonal streams arising from Chalk springs in response to seasonal water table fluctuations.
The relationship between surface karst features and hydrogeologically significant subsurface karstic
features in the Chalk is unclear. MacDonald et al. (1998) point out that there is an inherent bias in
the recorded solution features towards those that are easily visible from the ground surface such as
swallow holes or dolines, and not necessarily those that are hydrogeologically significant. Relatively
rare exposures of chalk faces at Quarries (e.g. Castle Lime Works Quarry, Hertfordshire) give some
indication of potentially more laterally persistent features. Solution enlarged fractures are observed
at Water End, Hertfordshire, which may indicate hydrogeolgical connections from the swallow holes
and dolines. The high concentration of swallow holes and dolines in the region of Water End may be
hydraulically significant as they are likely to provide a mechanism for directing a significant volume of
recharge into the chalk aquifer. In particular they may direct recharge rapidly to a depth at which more
laterally persistent solution features occur.
2.4.2 Evidence of rapid flow rates from tracer tests in the Chalk
Atkinson and Smith (1974) conducted a tracer test in the Hampshire Chalk, where swallow holes occur
near to the northern margins of the Eocene outcrop. Rhodamine WT dye was pumped into a swallow
hole for three days. The travel time to its emergence at the Bedhampton spring (a distance of 5.75 km)
was 62.5 hours, corresponding to a velocity5 of 2.2 km day−1 (peak concentration). Atkinson and Smith
(1974) concluded that turbulent flow in an open system of fissures widened by solution, or conduits, is
required to achieve the observed velocities.
Banks et al. (1995) conducted a tracer test in the Berkshire Chalk between the Holly Grove stream
sink and the Blue Pool spring 4.7 km away. Both the spring and the stream sink were less than 1 km
from the Chalk/Eocene boundary. The observed velocities were 5.8 km day−1 for peak concentration
and 6.8 km day−1 for breakthrough. The authors suggest that little attenuation occurred as the tracer
moved from the sinkhole to the spring. Maurice et al. (2006) carried out two injections of flourescein
tracer from the Smithcroft Copse (nearby to Holly Grove) to the Blue Pool. Maurice et al. (2006)
determined groundwater velocities of 5.12 and 4.71 km day−1 with tracer recoveries of 25.5 % and
21.7 % respectively.
In the Hertfordshire Chalk, there is abundant evidence of the existence of rapid preferential flow
routes within the Chalk. Water End is a well known site of swallow hole activity (Section 3.2.5). The
swallow holes are located close to boundary between the Chalk and Eocene cover. Three tracer tests,
carried out in 1927, 1928 and 1932 using fluorescein dye showed that water recharging the swallow holes
were detected in a series of springs and wells in the Lea Valley up to 16 km from the swallow holes
(Harold, 1937). The breakthrough times gave flow velocities up to 5.5 km day−1 , with tracers being
detected over a 30 ◦ arc from the swallow holes (Harold, 1937). The velocities between the swallow
holes and a particular detection point in the Lea Valley varied by up to 37 % between tests, and the
fastest route also varied between tests. A test carried out in 1935 in swallow holes at South Mimms
(3.5 km to the sourthwest of Water End) also showed flows towards the Lea Valley over distances up to
5 Velocities quoted are mimimum velocities calculated based on straight line distance between input and output locations
2.4. Karstic behaviour of the Chalk 36
• MS2 Coliphage were injected by borehole dilution in Harefield House Borehole, near Sandridge;
• Phi X174 phage were injected by borehole dilution at Comet Way Borehole, near Hatfield;
• Serratia Marcescens phage were added to a sinking stream close to a large swallow hole complex
at Water End near North Mymms.
Monitoring for the tracers was conducted at 21 locations throughout the study area, comprising abstrac-
tion wells, observation boreholes, springs and surface waters. The duration of the monitoring period was
approximately two months.
The connections indicated by the tracer tests are further described in Section 3.2.6. Results indicated
rapid groundwater flow between the Water End swallow holes and a wide spatial distribution of locations
in the Lea Valley. Rapid flow connections were also indicated between locations to the west of the main
karst system toward Hatfield, Sandridge and within the Vale of St. Albans.
Between Water End and the Lea Valley (Serratia Marcescens phage), travel times and distributions
were broadly consistent with the 1920s and 1930s tracing results. However, better data resolution al-
lowed for the observation of breakthrough curves and secondary tracer peaks suggesting multiple arrivals
of tracer. First detections indicated groundwater velocities of between 1.8 km day−1 and 3.9 km day−1 .
Overall tracer recovery was estimated at approximately 15 % of the injected tracer mass.
Identification of tracer breakthrough for the MS2 Coliphage and Phi X174 phage was complicated
by measured concentrations being close to background. However, observations suggest groundwater ve-
locities between Hatfield (Phi X174 phage, Comet Way BH) and Essendon PWS, Arkley Hole Spring and
Lynchmill Spring of between 0.8 km day−1 and 1.8 km day−1 . Groundwater velocities between San-
dridge (MS2 Coliphage, Harefield House BH) and Hatfield Quarry, Essendon PWS, Arkley Hole Spring
and Lynchmill Spring indicated groundwater velocities between 0.05 km day−1 and 0.9 km day−1 . In
both cases, attenuation was high: overall tracer recovery was estimated at less than 1 % of the injected
mass.
There are also examples of tracer tests indicative of karst flow in ares of the Chalk of Southern Eng-
land that are unconnected with obvious karstic features. Price et al. (1992) reported a tracer experiment
at the M1/M25 motorway intersection, an area close to the Chalk/Palaeogene boundary where karstic
2.4. Karstic behaviour of the Chalk 37
features are common, but the studies were carried out in soakaways that were apparently unassociated
with karstic features. Some tracer traveled rapidly to a pumping station a distance of 3 km away, with
recorded velocities in excess of 2.4 km day−1 . The tracer recovery was very low, and it was thought that
a significant fraction of the flow was moving through fine fractures.
A number of borehole dilution tests and radial tests from an injection borehole to a pumped bore-
hole, were carried out in East Anglia (Kachi, 1987; Ward, 1989). The main conclusion of the extensive
testing and modelling was that flow was dominated by micro-fractures with a range of sizes rather than
by a few discrete high-permeability conduits.
• Within the Newhaven, Lewes Nodular, New Pit, and Holywell Nodular Chalk members, conjugate
fractures are common. At the intersection of the fractures, cavities can develop which could give
rise to rapid groundwater flow and therefore high permeability.
• Marl horizons restrict vertical flow due to their high clay content. However, marl and flint bands
may act to ‘seed’ the development of larger solution features, e.g. by forcing groundwater moving
downwards from the Chalk above to dissipate horizontally.
• The presence of hardgrounds, where shallower than about 100 m below ground level, can sig-
nificantly increase the permeability of the Chalk. Hardgrounds (such as the Chalk Rock and the
Melbourne Rock) fracture more cleanly than other chalks due to their greater hardness (Price,
1987), and since the hardgrounds are better cemented, fractures within them tend to remain open
at greater depths than in softer chalk. The open fractures allow groundwater to flow through them
generating preferential flow paths which may then be enhanced by dissolution.
The structure of the Chalk can affect aquifer properties, although the relationship between structure
and permeability is complex. Generally, folding tends to increase the fracturing of the Chalk along the
axis of anticlines. Deep burial of the Chalk (e.g. beneath Palaeogene deposits of the London Basin)
tends to reduce permeability and storage . Rapid, anisotropic, groundwater flow is often associated with
fault zones in the Chalk (Allen et al., 1997): parallel to the fault groundwater flow is rapid, while it is
impeded perpendicular to the fault.
Price et al. (1993) reviewed the factors that may have contributed to the development high trans-
missivity along valleys. There may be a higher frequency of open fractures within valleys since valleys
often follow lines of structural weakness, and erosion along the valleys reduces effective stress, allowing
horizontal fractures to open. However, a higher frequency of fractures in not necessarily a prerequisite
to high permeability; it can develop by solution enhancement of a lower frequency of fractures. Price
(1987) illustrated how the increased flux of groundwater along a valley could give rise to high perme-
ability within the valleys and the development of a single zone of dissolution-enhanced fractures. The
2.5. Hierarchy in the Chalk aquifer 38
concentration of groundwater flux, and hence increase in velocity, occurs as groundwater flows from
recharge areas to discharge areas in the valleys. The mixing of groundwaters of different chemistry
towards discharge points can also increase the dissolution potential of groundwater.
Periglaciation is believed to have played an important role in the enhancement of permeability along
the valleys (Younger, 1989). Repeated freezing and thawing of the active layer would have broken down
the top few metres to a weathered chalk that would have been easily eroded. Furthermore, in the valleys,
the flow of surface water would have kept the ground unfrozen to a greater depth for large parts of the
year, forming a talik within the Chalk of the valley floor. Chalk is dissolved more easily under cold
conditions, therefore the concentrated flow of groundwater within these taliks and the low temperature
of the groundwater would combine to dissolve Chalk within factures. Within valleys, an increase of
fracturing is observed in the top 5–6 m of the Chalk. In some valleys, periglacial activity led to fractures
opening up to a depth of 20–30 m.
Geomorphological karstic features, and rapid groundwater flow, are generally more frequent close
to Tertiary cover (MacDonald et al., 1998). Based on a field survey of the Pang and Lambourn catchments
in Berkshire, Maurice et al. (2006) identified three distinctive geomorphic Zones characterised by the
density of surface karst features which was related to the proximity to the Palaeogene-Chalk contact.
Soils associated with Palaeogene deposits tend to be quite acidic. The soils associated with Quaternary
cover also tend to be clayey and therefore to concentrate runoff into discrete points. As recharge drains
through the cover, it remains unsaturated with respect to calcite until it reaches the Chalk surface, thus
allowing acidic recharge to be channelled to discrete points of the Chalk surface.
Solution pipes and swallow holes can allow acidic recharge to penetrate deep into the unsaturated
zone (and even the saturated zone). Therefore the acidic recharge can link up with fracture systems
within the aquifer and allow the enlargement of fractures to conduits. This is probably the mechanism
for the observed rapid groundwater flow near Quaternary cover (Allen et al., 1997). It is likely that zones
of slightly wider initial aperture occurring in fractures with asperities will be preferentially widened
by dissolution until discrete ‘conduits’ are formed (Younger and Elliot, 1995). Even once the cover
and geomorphological features have been moved by erosion, the deeper hydrogeological features will
remain, providing rapid groundwater flow and preferential flow paths.
Rapid groundwater flow and swallow holes are also often observed on valley bottoms, associated
with flowing streams. These ’karst’ type features may have a different origin to those observed on the
Chalk interfluves (Allen et al., 1997). Surface water flowing within streams has different chemistry
from the groundwater, therefore mixing can produce water that has a high dissolution potential. This
aggressive water, coupled with the high flux through the valleys and the dynamic of surface water flow
can produce sink holes and springs in stream beds, probably linked with conduits. Therefore, as in more
classic ‘karst’ environments, surface water can flow in discrete channels underground within valleys.
Figure 2.2: (a) An idealised double-porosity aquifer; (b) an idealised double-permeability aquifer. From
Price et al. (1993)
The porosity and permeability components contributed by the fissures are referred to as the as the
fissure (or fracture) porosity and permeability. The blocks bounded by the fissures are usually described
as matrix blocks, and the non-fissured fraction of the porosity and permeability as the matix (or matric)
porosity and permeability.
The behaviour of the Chalk is more complex than described in the classical double-porosity model.
The majority of the water within the matrix pore space does not represent usable groundwater storage;
effective groundwater storage is primarily within the fracture network and the larger pores (Section 2.6).
2.6. Porosity components of the Chalk aquifer 40
In terms of groundwater flow, there is a hierarchy in permeability components (Section 2.7) as a result of
solution enhancement of primary fractures to form fissures and karstic conduits. The Chalk has therefore
been described as exhibiting dual-permeability behaviour (Price et al., 1993).
The Chalk is increasingly recognised as possessing karstic characteristics in places (Section 2.4.2).
This additional flow component is neglected by classic double-porosity models (Section 6.2.6). More re-
cently, it has been proposed that carbonate aquifers, including the Chalk, be considered as triple porosity
and triple permeability aquifers (Worthington, 2003; White, 2003), to account for flow and storage in
three porosity and permeability elements:
was 0.0023, and the 25th and 75th percentiles were 0.0004 and 0.01 respectively. Approximately one
order of magnitude difference was recorded between estimates of storage coefficient from confined and
unconfined tests: unconfined tests had a median of 0.008 and confined tests had a median of 0.0006. The
authors suggest that the relatively small difference between unconfined and confined measurements may
be explained by the limited gravity drainage of the Chalk matrix, and therefore the relative importance
of elastic storage in both confined and unconfined conditions. However, MacDonald and Allen (2001)
also found a direct relationship between transmissivity and the storage coefficient, which indicates that
gravity drainage from the fracture network is still a significant component of storage.
In fissured rocks such as the Chalk, where nearly all groundwater movement occurs in the fissure
network, the porosity that is of importance when groundwater flow velocities are being considered is the
porosity that represents the saturated pore space which contributes to the flow. This porosity, which will
be practically equivalent to the porosity of the fractures, is significantly lower than the total porosity: in
general, the porosity of the fractures is around 0.01 % Price et al. (1993). De Marsily (1986) defines the
ratio of the volume of water able to circulate to the total volume of rock as the kinematic porosity. The
term effective porosity is also frequently used in this context. However, Ward et al. (1998) encourages the
use of the term kinematic porosity over effective porosity for the purposes of solute transport to prevent
association with the specific yield, which is sometimes taken to be representative of the effective porosity,
and has been used for the purposes of solute transport in fractures (e.g. Bibby (1981)). However, it may
not be appropriate to use values of specific yield in this context: specific yield values are usually derived
during hydraulic testing and the fluid volumes released due to the stresses applied during testing may be
unrepresentative of the volume of water through which solute transport occurs.
The effective porosity is related to the aperture of the active fractures and the frequency of occur-
rence of those fractures. Atkins (2004) estimate an effective porosity range of 0.05 % to 0.50 % for
the Hertfordshire Chalk, assuming fracture apertures between 0.5 mm and 5.0 mm and an average of
one fracture per metre of borehole. However, they note that this range does not take into account the
contribution of the vertical fractures to the effective porosity (vertical boreholes only detect horizontal
or sub-horizontal fractures) and therefore state a more ‘reasonable’ effective porosity range of 0.5 % to
2.0 %.
Values for effective porosity in a fractured rock can be determined if the Darcy velocity or flux q
and average linear velocity v are known through the relationship
q
ne = . (2.1)
v
Watson (2004) calculated effective porosity for the Chalk of the Tilmanstone-Eastry Valley in Kent
based on the results of single borehole dilution tests and natural gradient tracer tests. Single borehole
dilution tests were undertaken to obtain darcy velocity, q, for specific depth intervals. The groundwater
(fissure) velocity, v, was estimated from the arrival times of the natural gradient tracer tests. The effective
porosity was then calculated using equation 2.1. The calculated effective porosities were in the range
0.1 % to 0.3 %.
2.7. Permeability components of the Chalk aquifer 42
boreholes being drilled within the valleys, and the consequent lack of data available across the interfluves
hinders the understanding of the variations in transmissivity away from the valleys. However, Allen et al.
(1997) report that flow logging from boreholes drilled on interfluves of the Berkshire Downs indicated a
general thinning of the transmissive zone compared with valleys.
Transmissivity values were also found to be significantly higher in unconfined areas compared to
confined areas: unconfined sites had a median transmissivity of 920 m2 day−1 compared to a median
of 220 m2 day−1 at confined sites. It is thought that this reflects increased solution enhancement of
fractures under unconfined conditions.
Allen et al. (1997) speculated that a correlation between some function of the matrix pore size distri-
bution and hydraulic conductivity is likely to be more significant, although such a correlation was not
established for the Chalk matrix.
ga3
Tf = (2.2)
12ν
where g is the acceleration due to gravity and ν is the kinematic viscosity. The transmissivity of a
fracture is therefore very sensitive to aperture size since the transmissivity is a function of the cube of
the aperture. Barker (1993) highlights this by a simple example: equation 2.2 gives a transmissivity of
about 10−3 m2 s−1 for a fissure with an aperture of only 1 mm, and an enormous 1 m2 s−1 for a 1 cm
aperture. Equation 2.2 is valid providing flow is laminar (Section 2.8).
2.7. Permeability components of the Chalk aquifer 44
Figure 2.3: Relationsip between fissure spacing, aperture, porosity and hydraulic conductivity for a
fissure system containing three plane, parallel, mutually perpendicular smooth-walled fissures filled with
pure water at 10 degC and porosity relationship (Price et al., 1993).
Bloomfield (1996) measured fracture orientation, trace length, spacing and aperture using section
and scan-line surveys on the Upper Chalk at Play Hatch Quarry, Berkshire. There were two dominant
joint sets: a set parallel to bedding and a set at a high angle to bedding. The trace length and spacing
distributions of the two joint sets approximated to log-normal distributions, with geometric mean trace
lengths of 0.15 m and 0.30 m, and spacings of 0.10 m and 0.12 m, respectively. Calculated fracture
interconnectivity indices suggest that the bedding parallel joint set is likely to be of greater hydraulic
importance than the high-angle joint set. Bloomfield (1996) used the fracture interconnectivity index
of Rouleau and Gale (1985) to demonstrate that the bedding parallel joint set fractures had a higher
interconnectivity, suggesting that the are likely to be of greater hydraulic importance than the high-angle
2.7. Permeability components of the Chalk aquifer 45
joint set. Bloomfield (1996) proposed that the results of the Play Hatch Quarry support a visualisation of
the Chalk consisting of “scale-invariant fault-bounded segments, where the internal fracture architecture
of each segment is dominated by continuous bedding plane fractures, and subordinate, scale-dependent,
arrays of joints”. The scale of jointing within a given fault-bounded segment is a function of bedding
thickness.
Aperture measurements obtained for a single bedding plane fracture ranged from less than 0.5 mm
to 23.5 mm. Apertures approximated to a negative exponential distribution (with a mean of 1.2 mm)
below 7 mm, and to a log-normal distribution (with a mean of 11.7 mm) above 7 mm. It was inferred
that the larger apertures have been affected by solution processes and that flow through bedding plane
fractures is channeled across 10–20 % of the fracture surface area.
Watson (2004) calculated block sizes (fracture spacing) for the Chalk of the Tilmanstone-Eastry
valley in Kent corresponding to fracture density predicted by the relationship established for the area:
This relationship was determined based on data from scanline surveys in the area and a fracture profiled
produced with televiewer an optiviewer logging.
Watson (2004) used data for block sizes and effective porosity (Section 2.6) to calculate fracture
aperture using the relationship
In general, fracture apertures were found to increase with depth, and ranged from 0.44 mm to 3.83 mm.
Watson (2004) also calculated fracture apertures from hydraulic conductivity data from packer testing
using an approximation to the cubic law:
where N is the fracture density, a is the fracture aperture, and Kef f ective is the effective hydraulic con-
ductivity. Apertures determined by this method (cubic law or hydraulic apertures) ranged from 0.92 mm
to 1.39 mm.
Figure 2.4 summarises the range of fracture spacing and fracture aperture measurements from a
number of sources. It should be noted that fracture apertures are difficult to measure in-situ: at out-
crop, weathering effects may produce enlarged apertures, and accurate down-hole measurement is only
feasible for larger openings. As such, there are few direct observations for the Chalk, and apertures are
usually determined from cubic law approximations or in relation to fracture density and fracture porosity
measurements.
It has been shown that groundwater velocities measured from the tracer tests described in Sec-
tion 2.4.2 could result from flow through either small channel conduits or more laterally extensive fis-
sures. For the Hampshire study Atkinson and Smith (1974) calculated that, for the estimated head
gradient, the volume indicated by the tracer time to peak was equivalent to flow in a single circular pipe
2.7. Permeability components of the Chalk aquifer 46
Min
Fracture Spacing
Chalk (Kent) Max
Mean
Chalk (Yorkshire)
Near-surface (Oxfordshire)
0.01 0.1 1 10
Fracture spacing (m)
Fracture Aperture
Kent Chalk Min
Max
Yorkshire Chalk Mean
Yorkshire Chalk
Yorkshire Chalk
0.1 1 10 100
Fracture aperture (mm)
REFERENCES
Figure 2.4: Ranges of fracture spacings and fracture apertures for the Chalk. Results as cited in Bloom-
field (1996) and Watson (2004).
2.7. Permeability components of the Chalk aquifer 47
740 mm) in diameter (although in reality several features would probably be involved). Price (1987) sug-
gested that in an ideal case (a plane parallel fracture with no roughness or chanelling) a fissure of only
4.5 mm width, with a transmissivity of 5000 m2 day−1 , provides an alternative hydraulic explanation
for the observed combination of displacement volume and velocity. By using the same method Banks
et al. (1995) calculated that a single fissure 5.4 mm in width could theoretically be sufficient to represent
the fracture system in Berkshire, and similarly, Maurice et al. (2006) calculated that the observed veloc-
ities would suggest a comparable aperture of 4.9 mm. However, they note that the hydraulic gradient
(0.004) is calculated using the elevation difference between the sink and the spring, and since the water
table is lower than the surface elevation of the stream sink, the actual hydraulic gradient must be smaller,
and therefore the calculation must underestimate aperture. Cook (2010) determined apertures around
2 mm to 4 mm using the cubic law approach for the results of the tracer tests in Hertfordshire. These
results were also interpreted as minimum equivalent conduit diameters between 0.6 m and 2.1 m, based
on maximum recorded discharges and flow velocities at locations including Water End and Catherine
Bourne swallow holes, and Arkley Hole and Lynchmill Springs.
frequency and aperture of fractures declines due to increasing overburden and a general reduction in
circulating groundwater and hence dissolution. The greatest permeability in the Chalk is observed in
the zone of water table fluctuation where the movement of groundwater can enhance the aperture of the
fractures by dissolution. A similar pattern of permeability variation with depth is observed where the
Chalk is confined by younger deposits within the London and the Hampshire Basin. In the London Basin,
flow logging has illustrated that the majority of inflows to boreholes, shown by geophysical logging,
occur within 20 m to 30 m of the upper surface of the Chalk.
Williams et al. (2006) compared the results of the packer testing at Trumpletts Farm with results
from borehole imaging and geophysical testing; while some of the highly permeable zones appeared to
be associated with obvious fractures, large fractures could be seen in zones which had much lower per-
meability, and some highly permeable zones appeared to be associated with poorly developed fractures.
Therefore, not all fractures are hydraulically active. Furthermore, Williams et al. (2006) found differ-
ences in flow velocity depth profiles (from single borehole dilution tests) in the same borehole which
was tested both before and during a pumping test at an abstraction borehole about 40 m from the site,
indicating that different fractures become active when the aquifer is stressed. Therefore, field studies
show that flow near individual boreholes is highly heterogeneous, and that there is uncertainty in the re-
lationship between the characteristics of a fracture observed in a borehole and the amount of flow which
that fracture contributes to the borehole.
Groundwater flow in the Chalk is highly heterogeneous and borehole yields may vary by orders
of magnitude over distances of less that 100 m reflecting the complexity of flow in the aquifer. The
single borehole dilution testing by Williams et al. (2006) showed differences in flow velocity depth
profiles between boreholes located within a few tens of metres across the site. These are inferred because
the different boreholes, although of similar depth and drilled in very close proximity, intersect slightly
different parts of the fracture network an hence groundwater flow system. In particular, a flowing feature
at the base of one borehole is not intersected by the second, which is drilled from a slightly higher
elevation.
When considering flow observations in boreholes, it is important to note that the presence of the
borehole connects flowing horizons and allows vertical flow. Therefore, the observed flow may not be
representative of flow in the aquifer in the absence of the borehole.
the conduit system. Connectivity between cave conduits and fissures intercepted by quarrying has been
demonstrated in the highly karstic Carboniferous Limestone (Edwards et al., 1991). The occurrence
of turbidity and bacteria in some Chalk boreholes suggest that there may also be interaction between
fissures and larger scale conduits in the Chalk (MacDonald et al., 1998).
During the Water End tracer tests in Hertfordshire (Section 2.4.2), visible colouration demonstrated
connections to seven spring and borehole abstraction sites up to 6 km apart. Visible colouration in the
Blue Pool complex during the tracer tests in Berkshire by Maurice et al. (2006) demonstrated that tracer
was discharged at a number of sites up to 100 m apart, suggesting flow through laterally extensive fissures
in the vicinity of the springs. The results of both these tracer studies are consistent with the downstream
sections of conduit systems being characterised by fissure flow resulting in lateral dispersion across a
large scale three-dimensional network.
Significant loss of tracer mass is indicated by the recovery data from the Hampshire and Berkshire
tracer tests: tracer recoveries were 70 % (Atkinson and Smith, 1974) and 25 % and 21 % (Maurice et al.,
2006) respectively. The breakthrough curves displayed a steep falling limb after the main peak, followed
by a long flat tail of low concentration. Such a pattern is indicative of exchange between mobile water
and immobile water by the mechanism of double porosity diffusion (Section 2.9.4), but the rapid travel
time implies that for a single conduit or fissure, tracer transport would be too rapid for double porosity
diffusion to account for the tracer loss. Maurice et al. (2006) invokes a complex flowpath, with sections
of the flowpath characterised by multiple pathways with dispersion from the main conduit into smaller
fissures and fractures, as a possible explanation for the loss of tracer.
The tracer test from soakaways at Bricket Wood in Hertfordshire (Price et al., 1992) showed very
low tracer recoveries of <0.01 % which the authors propose reflect a significant fraction of the flow
occurring in fractures and fissures with relatively small apertures, allowing attenuation of the tracer by
diffusion into the Chalk matrix.
where ν is the mean velocity of a fluid with density ρ and dynamic viscosity µ passing through a pipe of
diameter d. In a porous or fissured medium, d becomes a representative length dimension characterising
interstitial pore-space diameter or fissure width (Ford and Williams, 2007). Laminar flow generally
occurs for Re <2300.
For flow in pipes, under laminar conditions, specific discharge can be evaluated by what is termed
Poiseuille’s law:
πd4 ρg dh
q= . (2.5)
128µ dl
Increasing velocity, sinuosity and roughness may eventually result in flow through the tube be-
coming turbulent. For turbulent flow, the specific discharge can be calculated by the Darcy-Weisbach
equation:
2dg dh
q2 = . (2.6)
f dl
where f is a friction factor.
Turbulent flow conditions frequently arise in pipes and fissures in karst (Ford and Williams, 2007).
In karst, the range of conditions under which Darcy’s Law can be considered valid is very restricted: it
only applies in conditions that permit velocities to be low, and this usually involves some combination
of relatively small aperture and low hydraulic gradient (Ford and Williams, 2007).
For the range of karstic flow velocities observed in the Hertfordshire tracer tests (0.022–
0.068 m s−1 ), turbulent flows (where Re >2300) would occur above a conduit diameter of 0.04–0.13 m
(Cook, 2010). Therefore, the majority of karst flows within the Hertfordshire karst conduits are expected
to be turbulent.
2.9.1 Advection
Advection is the term used to describe movement attributed to transport by the flowing groundwater:
advecting solutes travel at the same rate as the average linear velocity of the groundwater.
The groundwater velocity, v, is given by the volumetric flux, q, divided by the kinematic porosity,
ne :
q
v=
ne
The volumetric flux, q, can be determined from Darcy’s law (equation 2.3).
2.9.2 Adsorption
Solutes can be attenuated relative to advective transport by the process of adsorption. Most of the sur-
face area of chalk onto which adsorption can take place is within the rock matrix. Mineral or organic
deposits are quite often observed on the surfaces of fissures, and when present, must have a significant
2.9. Solute transport in the saturated zone of the Chalk 51
impact on the amount of adsorption taking place, by providing adsorption sites and acting as a barrier
for diffusion into the matrix (Barker, 1993). Some models include a fracture skin to take some account
of the phenomenon.
2.9.3 Dispersion
Dispersion refers to the process of spreading (of solutes, particles or heat) during transportation. Ma-
trix diffusion and adsorption both have dispersive effects. The term mechanical dispersion refers more
specifically to the spreading caused by variations in groundwater flow velocity. Dispersion arises because
of the detailed variations in flow velocity in pores and fractures mainly due to the complex splitting and
joining of paths but also due to flow velocity variations in single paths. Strictly speaking, mechanical
dispersion is not a process in its own right; it is rather an expression of the fine (often random) detail
of the advection process. Therefore, if advective transport could be fully characterised throughout the
system, there would be no additional dispersion phenomenon to consider. Dispersion must therefore be
related to the advective model (conceptual or mathematical) in use, particularly to its scale of averaging
(Ward et al., 1998).
The normal approach to describing transport in a dispersive medium is via the convection-dispersion
equation (Equation 2.7 for flow in one-dimension), which contains a characteristic dispersion coefficient,
D.
∂C ∂2C ∂C
=D 2 −v (2.7)
∂t ∂x ∂x
The dispersion coefficient is normally considered to increase in proportion to the absolute value of
the velocity, v, so D = αv where α is known as the dispersivity. Dispersion is sometimes separated into
longitudinal and transverse dispersion to refer, respectively, to dispersion in the direction of flow and
dispersion perpendicular to that direction. Transverse dispersivity is typically much less than longitudinal
dispersivity. Dispersion coefficients for single fractures are more likely to show proportionality to the
square of velocity rather than to the velocity (Ward et al., 1998).
In practice, the processes of molecular diffusion and mechanical dispersivity cannot be separated in
flowing groundwater. Instead, a factor termed the coefficient of hydrodynamic dispersion is introduced
which takes into account both mechanical mixing and diffusion:
D = αv + D∗
where D is the coefficient of hydrodynamic dispersion, D∗ the effective molecular diffusion coefficient.
An analytical solution to Equation 2.7 was provided by Ogata and Banks (1961).
Dispersion in individual fissures is likely to be small and dependent on the fissure aperture, rough-
ness etc., and at a larger scale dispersion depends on the interconnecivity of fractures (Grisak and Pick-
ens, 1980, 1981). Barker (1993) demonstrated that dispersion, although not negligible, can normally be
ignored, in relation to matrix diffusion, when modelling solute transport in the Chalk.
2.9. Solute transport in the saturated zone of the Chalk 52
2.9.4 Diffusion
Molecular diffusion represents the net movement of solute under a concentration gradient and can be
described by Fick’s first and second laws. Molecular diffusion of contaminants is not normally of prac-
tical consideration where advection and mechanical dispersion are dominant. However, within the water
in the chalk matrix, which is (effectively) immobile, transport of solutes can take place by molecular
diffusion. For any diffusive process, a characteristic time for diffusion over a distance x can be defined
x2
as D, where D is the diffusivity (Diffusion coefficient divided by porosity) (Barker, 1993).
For the double-porosity Chalk, and other comparable fractured porous media, Foster (1975) pro-
posed that a major component of solute movement was controlled by a mechanism involving solute
exchange, through lateral molecular diffusion, between mobile fissure water and (relatively) immobile
matrix water. This double-porosity diffusive exchange has been demonstrated to be of significance to
the interpretation of thermonuclear tritium and pollutants, such as nitrate, in the unsaturated zone of
the Chalk (Foster, 1975; Barker and Foster, 1981) and when predicting the rate of lateral migration of
pollutants in the saturated zone of the aquifer (Watson, 2004; Burgess et al., 2005).
Double-porosity diffusive exchange acts to attenuate contaminants and significantly prolongs the
duration of contamination (Figure 2.5). Considering an initially contaminated fracture water, and un-
contaminated matrix water, there will be a diffusive flux of the dissolved contaminant from the fracture
water into the matrix water. The movement of the contaminant down hydraulic gradient will therefore be
retarded compared to transport by advection. At a later stage, when the primary source of contamination
input to the fracture water has ceased, there will be a diffusive flux from the now contaminated matrix
water to the fracture water. The matrix water therefore acts as a secondary source of contaminant input,
which prolongs the duration of contamination detected down-gradient.
∂c ∂2c
= DA 2 (2.9)
∂t ∂x
DE = φD DA (2.10)
Figure 2.5: Double-porosity diffusive exchange of solutes. At an early stage, diffusion from the contam-
inated fracture water into the matrix water acts to retard the transport of contaminants down-gradient. At
a later stage, contaminated porewater acts as a persistent secondary source of contamination.
2.10. Flow and transport in the unsaturated zone of the Chalk 54
using tritium) in samples of both fissured and unfissured chalk (Table 2.1). These values represent the
mass flux through the saturated matrix per unit concentration gradient in the water. Hill’s values are
appropriate for use in Fick’s first law of diffusion. In a critical discussion of Hill’s results, Muller (1987
deduced that the ratio of the diffusion coefficient to the free water diffusion coefficients is about 0.25.
So the diffusion coefficient in chalk can be estimated as one quarter of the free water value, if this value
is known for a solute.
Table 2.1: Molecular diffusion coefficients in fissured and unfissured chalk. After Hill (1984). These
values represent the mass flux through the saturated matrix per unit concentration gradient in the water.
No values of diffusion coefficients for bromate or bromide have been reported in the literature. In
the absence of a specific value for bromide and bromate, it seems reasonable to expect the bromide and
bromate ions to behave similarly to chloride and nitrate ions.
flow could be reconciled with a much slower observed downward movement of tritium: fissures within
the Chalk would focus tritium input to the unsaturated zone and the concentration gradient between the
contaminated fracture water and the matrix water would cause lateral diffusion of the solute into the ma-
trix (i.e. double-porosity diffusive exchange discussed in Section 2.9.4), greatly retarding its downward
movement. Simulations using a double porosity diffusion exchange model (Barker and Foster, 1981)
indicated that this mechanism lead to preservation of the tritium profile in the unsaturated zone with only
minor dispersion. However, Mathias et al. (2005) showed that double porosity models which assume
matrix flow to be negligible (e.g. Barker and Foster 1981), require an unrealistically small fracture spac-
ing (<25 cm) to preserve peaks without ‘solute spreading’ through the profile. Mathias et al. (2005)
analysed the impact of flow in the matrix by comparing a double porosity model (based on Barker 1982)
with an equivalent double-permeability model with a portion of flow in the matrix and demonstrated that
solute spreading in such models can only be reduced (whilst using sensible estimates of fracture spacing
and diffusion coefficients) by allowing for a portion of flow in the matrix. Thus, Mathias et al. (2005)
argues that flow in the matrix of the Chalk of the unsaturated zone is significant.
However, Mathias et al. (2005) was not able to obtain a sensible estimate for the proportion of total
infiltration that enters the matrix as the model assumed steady-state flow conditions necessitating the use
of annual mean estimates of infiltration. The assumption of steady-state flow also forces fracture flow
to be either negligible or persistent whereas in reality it is likely to be intermittent (Price et al., 2000).
Subsequently, Mathias et al. (2006) developed a transient one-dimensional double permeability model
for the unsaturated zone of the Chalk. This model indicated that infiltration (as calculated by simple
two-store models) needs to be significantly attenuated to ensure that enough flow occurs in the matrix
such that solute spreading is reduced to a reasonable level. The justification for the attenuation was the
existence of soil and gravelly chalk layers. Mathias et al. (2006) demonstrated that such a model was
compatible with a fast water table response: there was a time lag of only three days between effective
precipitation input and water table flux.
1. the continuum approach, which assumes that the fractures mass is hydraulically equivalent to a
porous, granular medium, i.e. an equivalent porous medium (EPM) model; and
2. the discontinuum or discrete approach, which assumes that the rock cannot be characterised as a
granular medium, and so considers that flow is best dealt with in individual fractures or fracture
sets.
The appropriate model to simulate transport of water and/or solutes within fissured systems such as
the Chalk depends on a consideration of how the behaviour of a fissured system is related to the time-
2.11. Modelling flow and transport in fissured rocks 56
scales of the transport processes. For double-porosity systems, with advective transport in the fissures
and diffusive transport in the matrix, the suitable model representation depends on the time-scale of the
process under consideration in relation to the characteristic times for diffusion across a fissure or a matrix
block (Barker, 1993).
1. When the fissures act independently of the matrix. If the time-scale of interest is small with
respect to the characteristic time for diffusion across the fissure width, then the effects of the
porous matrix can be ignored (because of both the restricted diffusion out of the fissure and of the
small volume of matrix accessed, in relation to fissure volume). Under these conditions (which
rarely exist outside a laboratory), the chalk can be modelled with an EPM model with a porosity
equal to the fissure porosity.
2. When the fissures and matrix act in unison. If the time taken for diffusive equilibrium between
fissures and matrix is small in relation to the time for any significant change in the fissure system,
then the chalk will behave as a (locally) homogeneous medium characterised by the total porosity.
A Quasi-steady-state (QSS) double-porosity model might also be adequate (Section 2.11.2.2).
EPM models are commonly used for regional water resources models, where the fissured system is
represented as homogeneous, with storage and permeability parameters characteristic of the matrix and
fractures combined.
conditions, Barker (1993) considers that a general diffusive-type DP model should be used, (although a
QSS-type DP model may be adequate over some periods).
2.11.2.3 Importance of matrix diffusion for water and solute transport in the Chalk
Characteristic times for hydraulic diffusion in the chalk matrix are around 5–500 seconds and for solute
diffusion in the chalk matrix are 50–500 years (Barker, 1993). Therefore, for water transport, matrix
diffusion will only be significant for very rapidly changing conditions, and it is reasonable to adopt a
QSS model for all but the most rapid transient pumping tests. In contrast, matrix diffusion will have an
important effect on solute transport over most time-scales of interest for contamination incidents, and
diffusive-type double-porosity models should be adopted.
solute concentrations in the fissure and matrix water. Numerical inversion of the transforms was then
used to investigate characteristic behaviour of the model for a number of special cases.
The porosity ratio is also related to the diffusion porosity φD , for example for a simple slab model:
2bφD
σ= (2.11)
a
DE
The diffusion porosity is defined as φD = DA , which is the ratio of the ‘apparent diffusion coeffi-
cient’, DA , (as appears in Fick’s second law) and the ‘effective diffusion coefficient’, DE , (as appears
in Fick’s first law). This porosity has been referred to in the literature as a fictitious porosity and can be
somewhat less than the total porosity (Barker et al., 2000).
6 In molecular diffusion, the mean-square distance traversed by a particle in time t is given by 2Dt, where D is the diffusion
coefficient.
2.12. Diffusion exchange model for solute transport in fissured porous rocks 59
Figure 2.6: Governing equations and assumptions for a double-porosity mocel with slab geometry. After
Barker (1982).
2.13. Summary 60
`2
tcb = (2.12)
DA
Table 2.2: Characteristic times for infinite slab geometry, with slabs of thickness 2b separated by frac-
tures of aperture a. For this model, the ratio of volume to area for a matrix block (`) is represented by
b.
x
ta = v Advection time in mobile phase
2
b
tcb = Dim Characteristic time for diffusion across a matrix block
tcf = tσcb2 Characteristic time for diffusion from a fracture into an equal volume of matrix water
σ = θθim
m
The ratio of matrix to fracture porosity
2.13 Summary
The behaviour of the Chalk as an aquifer is complex, and results from a combination of porosity and
permeability components that are a consequence of the Chalk lithology, tectonic history and weathering
and erosional processes. The Chalk is composed of very fine grained calcium carbonate micro-fossil
fragments which form a a highly porous, yet essentially impermeable, matrix. More than 95 % of
water in the Chalk is held in the interstices of the rock matrix, but the pore spaces are so small that
this water is effectively immobile. The mobile water (the remaining 5 % or less) is held within the
2.13. Summary 61
fractures that transect the chalk matrix. Some fractures have been enlarged by dissolution to become
fissures or even karstic conduits. The fissures and conduits provide the permeable pathways for flow.
Within the unsaturated zone of the Chalk, although the dominant flow pathways are via the fissures, a
small but significinat portion of flow is thought to occur within the matrix (Mathias et al., 2005, 2006).
These multiple components of porosity and permeability within the Chalk have long been recognised,
and it has been described as a double-porosity (dual-porosity) aquifer (Foster, 1975; Price, 1987; Barker,
1991; Price et al., 1993), a double-permeability (dual-permeability) aquifer (Price et al., 1993), and a
triple-porosity and/or triple-permeability aquifer (Worthington, 2003; White, 2003).
The Chalk is increasingly recognised as possessing karstic characteristics. In the Hertfordshire
Chalk, there is abundant evidence of the existence of rapid preferential flow routes within the Chalk.
Swallow holes and other dissolution features tend to be located close to boundary between the Chalk and
Eocene cover, and are particularly concentrated in the Water End area. Tracer tests have shown that rapid
groundwater flow occurs between swallow holes and stream sinks in the Water End area and springs and
boreholes in the Lea Valley, which is indicative of a dispersive system of karstic conduits.
In a multiple-porosity aquifer such as the Chalk, solute transport is dominated by two processes:
advection in fissures and diffusional exchange of solutes between fissures and matrix porewater. Adsorp-
tion may also affect the transport of some solutes. At larger scales, the effects of dispersion across the
network of fissures may become important.
Double-porosity diffusive exchange of of solutes between fissures and matrix porewater acts to
attenuate contaminants and significantly prolongs the duration of contamination. Considering an initially
contaminated fracture water, and uncontaminated matrix water, there will be a diffusive flux of the
dissolved contaminant from the fracture water into the matrix water. The movement of the contaminant
down hydraulic gradient will therefore be retarded compared to transport by advection. At a later stage,
when the primary source of contamination input to the fracture water has ceased, there will be a diffusive
flux from the now contaminated matrix water to the fracture water. The matrix water therefore acts as
a secondary source of contaminant input, which prolongs the duration of contamination detected down-
gradient. Double-porosity diffusion between mobile fissure water and immobile matrix water can be
described mathematically using Fick’s Laws of diffusion (e.g. Barker and Foster, 1981; Barker, 1982,
1985b), and has been demonstrated to be of significance when predicting the rate of lateral migration of
pollutants in the saturated zone of the aquifer (Watson, 2004; Burgess et al., 2005).
62
Chapter 3
3.2.2 Hydrology
River flows, rainfall and potential evapotranspiration data are reviewed and analysed in Buckle (2002)
and Atkins (2004). Much of the data were collated in work by Entec (2000) in connection with the Upper
j
k
!
A
Ri
± h
WARE r As
Riv ve
River Beane
! Ri BROADMEADS
Ri
A er
! ve
v e r R ib
A rL Mi
mr ! !
HARPENDEN am Y
X Y
XA A AMWELL
ea WELWYN
(o GARDEN HERTFORD ChadwellN HILL Amwell
rL !
Spring e A X
w
ee CITY Y !Spring
AAMWELL
) Y
X
MARSH
R iv
er
k
j !
A
Arkley Hole RYE COMMON
Rye House
SANDRIDGE Spring j
k
ESSENDON YSpring
! Y MIDDLEFIELD RDX
HATFIELD A kj X ! !
QUARRY FORMER
A A
!
A j
k
j WATERHALL
k !
AERODROME HODDESDON A
k
jk j
k k
j j
k j
k YLynchmill
X
HATFIELDk j j
jjk
k jk
kkj jkj
j kj k
k
jk
jj Spring
k
!
A jk
k
j
k j
k
! k
j
Ri
s
ve A j
k !
A
ro o
rV ST ALBANS
an
BROXBOURNE
nb
er
!
A Y
X
lb
El l e
A
! !
A
t
A ROESTOCK
Ground Elevation
TYTTENHANGER
fS
m OD Water
j End
k kk
jjk
j
o j
kj
k !
10 - 20 j
k
NORTHk
j
j
k jkj A
k
jMYMMS Cuk
r
21 - 30
l eLONDON e
j j
k
j
k
j
k
A!k ffl
e j
k TURNFORD
Va COLNEY oln j
k j
jk
jk
k
y
31 - 40 C j
k
j
k j
k
j
k
Ri v e
r
41 - 50 Ri v e jk
jk j
Br o
j
k k k
j jk
kj k
j
j
k
Rive r Le a (or L
51 - 60
Ne w
ok
61 - 70 NETHERWILD k
jBourne k j
ee)
! k
j j j j
k
A jk
k i ne k jk
jk
71 - 80 er POTTERS
SOUTH
3.2. Geology and Hydrogeology of Hertfordshire
th
81 - 90 SHENLEY MIMMS BAR Legend
Ca
91 - 100 RADLETT Springs
Y
X
Mymmshall Brook
101 - 110
k
j Swallow Holes
111 - 120
!
A Public Water Supply
121 - 130
rivers and streams
131 - 140
141 - 150 lakes
0 1 2 4 6 8 10 !
© Crown Copyright/database right 2008. An Ordnance A
Survey/EDINA supplied service.
Kilometers Geological Map Data © NERC 2008.
63
Mimram study, and along with additional data available from the Environment Agency.
In the east of the area, the upper River Lea and the River Ver flow south-east from the Chiltern
Hills. The surface water divide between the Lee and the Ver catchments is not well defined in terms
of topography, but it runs approximately south-east through Harpenden and Hatfield and also forms the
surface water divide between the catchments of the River Lee and the River Colne (Figure 3.1). The
River Colne flows southwest along the foot of the Palaeogene escarpment in the Vale of St Albans to join
the River Ver in the south-west of the study area. Significant groundwater–surface water interactions
occur in both the River Lee and River Colne catchments (Section 3.2.7).
The middle River Lee flows east through the central part of the study area along the northern foot
of the Palaeogene escarpment. Downstream of Hatfield, the Lee swings north-east and is joined near
Hertford by the Rivers Mimram, Beane and Rib flowing from the Chalk upland to the north, and further
downstream by the Rivers Ash and Stort. The River Lee then swings to flow south towards the River
Thames. Along this southerly flowing section, the middle and lower Lee is joined by a number of rivers
draining the Palaeogene escarpment. South of Hertford, the New River (an aqueduct constructed in the
17th Century) runs to the west of, and parallel to, the River Lea. The New River is fed from the River
Lee upstream of Ware and also accepts discharge from the Chadwell Spring when the spring is flowing.
The New River also takes pumped discharge from pumping stations of the Northern New River well
field.
In the west of the study area, the Catherine Bourne and the Mimmshall Brook drain to the River
Colne from the west of the Palaeogene escarpment. For most of the year, the Mimmshall Brook and
Catherine Bourne drain to a series of stream sinks and swallow holes near Water End. At times of
overflow, a spillway carries water north-west to the River Colne.
3.2.3 Geology
3.2.3.2 Lithostratigraphy
The generalised lithostratigraphy of the Hertford district is summarised in Table 3.1.
Ri
Solid Geology Riv
LONDON CLAY FORMATION
er
Mi
Chadwell h
Ri mr Spring r As
LAMBETH GROUP
ve am ve ±
v e r R ib
rL Ri
LEWES NODULAR CHALK Fm ea Y
X Y
X
SEAFORD CHALK Fm
River Beane
NEWHAVEN CHALK Fm
(o
rL
(UNDIFFERENTIATED) ee
) Y
X
Y
X Ne
w
R
Spring j
k
Y
X Y
X
k
j
j
k j
k
j
kk k j
k j
k j
k Y
X
j k
jjk
k k
jk
jjj k
j j
k
k
j j
k
s
k
k
jj
k
jk
jkj kj
j
k k
j
an
r
lb
Y
X
Ri v e
tA
S
Ne w
of
Water
l e End
Va j
k
j
k kk
jjk
j
e j
k
k
jj
k
j
k jk
k j
ln k
jk j
k
Co j j
r
j
k jk
k j
k
Ri v e j
k j
k j j
jk
kk
j
k j
k
j
k
3.2. Geology and Hydrogeology of Hertfordshire
jk
jk j
k k
j jk
k
Riv e r L ea
j
k j
kjkj
j
k j
k
j j
k ne k j j
k Legend
kk
j o ur jk
jk
B Springs
e
Y
X
rin
j
k Swallows
rivers
Cathe
Mymm shal l Brook
0 1 2 4 6 8 10 © Crown Copyright/database right 2008. An Ordnance Survey/EDINA supplied service.
Kilometers Geological Map Data © NERC 2008.
65
v e r R ib
rL Ri
ea Y
X Y
X
River Beane
(o
rL
ee
) Y
X
Y
X Ne
w
R
SITE Spring j
k
Y
X Y
X
k
j
j
k j
k
j
kk k j
k j
k j
k Y
X
j k
jjk
k k
jk
jjj k
j j
k
k
j j
k
s
k
k
jj
k
jk
jkj kj
j
k k
j
an
r
lb
Y
X
Ri v e
tA
S
Ne w
of
Water
l e End
Va j
k
j
k kk
jjk
j
e j
k
k
jj
k
j
k jk
k j
ln k
jk j
k
Co j j
r
j
k jk
k j
k Drift Geology
Ri v e j
k j
k j j
jk
kk
j
k j
k
j
k Alluvium
3.2. Geology and Hydrogeology of Hertfordshire
jk
jk j Valley Gravel
k k
j jk
k
Riv e r L ea
j
k j
kjkj Brickearth
j
k j
k Solid Geology Sand & Gravel
j j
k
j k ur
ne kjkj
jk Legend Taplow Gravel
jk
k o LONDON CLAY FORMATION
B Glacial Gravel
e
Springs LAMBETH GROUP
Y
X Boulder Clay
rin
LEWES NODULAR CHALK Fm
j
k Swallows SEAFORD CHALK Fm Clay-with-flints
NEWHAVEN CHALK Fm
rivers (UNDIFFERENTIATED) Pebble Gravel
Cathe
Mymm shal l Brook
0 1 2 4 6 8 10 © Crown Copyright/database right 2008. An Ordnance Survey/EDINA supplied service.
Kilometers Geological Map Data © NERC 2008.
66
Table 3.1: Lithostratigraphy of the Hertford district. After Bloomfield et al. (2004).
3.2. Geology and Hydrogeology of Hertfordshire 68
Table 3.2: Lithostratigraphy of the Chalk of the Hertford district. Based on Woods (2003).
3.2. Geology and Hydrogeology of Hertfordshire 69
Woods (2003) provided a detailed description of the various Chalk formations in the Hertford Dis-
trict (BGS Sheet 239), subdividing the Chalk Group into 5 lithostratigraphical units on the basis of
borehole resistivity and gamma log interpretations (Table 3.2). The lithostratigraphical units correspond
to the revised Chalk Group stratigraphy following Rawson et al. (2001) (Section 2.2.1). Cross-sections
produced from the resistivity log interpretations show the general southeasterly dip of the individual
Chalk subgroups. The New Pit Chalk outcrops in the far west of the study area. Moving east, the Lewes
Nodular Chalk is at outcrop up to Hatfield, and then the Seaford Chalk outcrops to the east of Hatfield.
Borehole correlations suggest that there are no major lateral changes in the development of Chalk Group
lithostratigraphical units across the district, except for slight thinning of the New Pit Chalk and a slight
expansion of the Lewes Nodular Chalk (Woods, 2003).
3.2.4 Hydrogeology
Gravels) in hydraulic continuity with the Chalk, and an upper perched sand and gravel aquifer above the
Boulder Clay. The degree of continuity between the upper sand and gravel aquifer and the Chalk aquifer
system is dependent on the extent and thickness of clay layers, which impede vertical flow. In places,
the Chalk-PTG aquifer system is only partially saturated and unconfined, elsewhere it is semi-confined
to confined by Boulder Clay.
In the southeast of the study area the Chalk is overlain by the Palaeogene Deposits, comprising the
Reading Formation of the Lambeth Group (formerly Woolwich and Reading Beds) and the London Clay.
The Thanet Sands are absent in the area. The Reading Fm sediments are unsaturated over the majority of
the study area, although in the far southeast of the study area they become saturated, and together with
the Chalk are confined by the London Clay aquiclude.
From a hydrogeological point of view, a unit known as the ‘Basal Sands’ is used to describe the
Palaeogene sediments that are in hydraulic continuity with the Chalk aquifer of the London Basin. In
general the ‘Basal Sands’ comprises the Thanet Sand and the lower part of the Lambeth Group. The top
of this unit is defined non-stratigraphically as the lowest clay greater than 3 m thick in the Palaeogene
succession (Board, 1972). In the study area, the London Clay aquiclude acts as the confining layer to the
Chalk-Basal Sands aquifer unit (Buckle, 2002).
1000 m2 day−1 to 4000 m2 day−1 . This equates to a hydraulic conductivity values between 3 m day−1
and 30 m day−1 .
Table 3.4: Chalk Group Aquifer Potential. After Mortimore et al. (1990).
are more sandy, recharge may occur directly into the Chalk without the generation of surface streams
(although the density and size of dissolution pipes developed beneath the cover may be greater). Most
stream sinks are fed by small, usually ephemeral, streams, although larger perennial streams occur. The
sinkholes/swallow holes allow rapid percolation of surface water to the Chalk water table.
The greatest concentration of stream sinks is in the North Mimms–Water End area. The behaviour
of the swallow holes in the North Mimms/Water End area has been described by Whittaker (1921),
Walsh and Ockenden (1982), Harold (1937). The area is part of the surface water catchment of the
Colne, and includes the Mymmshall Brook, its tributary the Catherine Bourne, and the Welham Green
Brook, which meet at Water End. The upper streams rise in the Palaegene escarpment, where they are
underlain by London Clay, and flow northwards. The streams disappear underground in a number of
sinks within the lower parts of the main valley and tributary valleys, which are underlain by Lambeth
Group deposits (Reading Fm) and/or Chalk. The swallow holes that appear to receive the greatest volume
of water are located at Water End (Walsh and Ockenden, 1982). The actual sinks used by the streams
depend on the flow (Walsh and Ockenden, 1982). During dry periods, the feeder stream sinks at several
points before reaching the main sinkhole complex. During periods of wet weather, flows reach the Water
End swallow holes and a lake forms in the depression occupied by the swallow holes. When the capacity
is exceeded, the area overflows to the upper reaches of the River Colne via a (normally dry) channel
which flows beneath the A1(M) which runs north-south through Hatfield (Figure 1.1).
A second major set of swallow holes occurs in the Hatfield area (Whittaker, 1921). It is thought
that that it drains east to the large springs at Arkley Hole, and may be linked to the conduit system
draining the Water End sinks. Also, numerous swallow holes (sinks or dolines) were mapped by Price
et al. (1989) near the M25/M1 junction (Figure 1.1) at Bricket Wood in the far west of the study area
. Many of the sinks are associated with the margins of dry valleys and were interpreted by Price et al.
(1989) as having been caused by groundwater issuing from the Chalk during the Pleistocene period under
periglacial conditions.
The major springs, which may be fed by karstic conduits, are the Chadwell Spring, and the Arkley
Hole Spring. Whittaker (1921) described several other springs between Amwell and Rye House, and
between Hoddesdon and Broxbourne. Chadwell Spring, a large spring just west of Hertford, is thought
to act as an estavelle (Whittaker, 1921): it runs turbid after heavy rain indicating conduit flow, but in
drier weather, it ceases to flow and takes water from the New River. Bloomfield et al. (2004) report that
it is suspected that the Arkley Hole springs are fed by the sinks in the Hatfield area and possibly the
Palaeogene outliers to the north, although it is not clear what evidence this is based on.
flow occurs is widespread. Furthermore, the series of tests show that the individual pathlines followed
by flow are not constant and may vary with water level. Cook (2010) considers that the major conduits
in Hertfordshire are most likely to be developed at or just below the zone of water table fluctuation. This
interpretation is supported by relatively stable groundwater elevations east of Hatfield (Section 3.3): the
water table may be controlled by the high transmissivity of such features.
Cook (2010) integrated the information from the historic and recent tracer tests, in combination
with a consideration of structural controls on groundwater flow, to form a new quantitative conceptual
understanding of the function of the karstic flow system in Hertfordshire. The tracer tests undertaken
in the study area suggest that there is a distributive karst flow system in Hertfordshire developed in
a broadly north-east direction between North Mymms and the Lee Valley (Figure 3.28). Tracing from
Water End and the Catherine Bourne has provided evidence that connectivity to the karstic features of the
Mymmshall Brook system extends along the entire Palaeocene feather edge between the Lee Valley as
far south as Turnford PS and at least as far west as south east Hatfield and also possibly to north-western
Hatfield. The overall system appears to comprise a recharge area comprising a convergent network of
conduits centered around the Mymmshall Brook Catchment and the North Mymms water table mound.
This then drains via a solution-enhanced pathway adjacent to the Chalk-Palaeocene boundary and via a
distributive network to springs in the Lee valley.
This distributive karst flow system is characterised by rapid, low attenuation transport. Cook (2010)
interprets the progression of tracer arrivals to suggest that flow paths could be coincident with the pattern
of surface karst and swallow holes between the feather edge of the Palaeocene outcrop and the River
Lee. This provides an alternative interpretation to earlier conceptual models, e.g. Buckle (2002), which
suggested a fan like series of major flow routes between Water End and the Lee Valley. Cook (2010)
proposes that the pattern of tracer breakthroughs could suggest that the Northern Loop of the Palaeocene
Feather Edge and River Lee is short-cut by the subsurface karst system, with a more direct karst com-
ponent to the central Lee Valley, and a dispersive component beyond Arkley Hole to the northern Lee
Valley.
Cook (2010) takes the consistency of connections for all three tracers to major springs at Arkley
Hole and Lynchmill Spring to suggest that these springs terminate karstic flow routes established prior
to the more recent development of abstraction wells. Furthermore, recovery of all three tracer species
at Essendon PWS as well as similarities with both bromate concentrations and turbidity at Arkley Hole
implies that this groundwater source is also probably directly connected to the same system. Therefore,
the implication is that a rapid approximately east-west aligned major flow pathway exists between Es-
sendon and the Southern Lee Valley, principally to the Lynchmill Spring, but also a likely distributary
connection to Amwell Marsh, the Rye House/Rye Common Area, Hoddesdon and further south to Turn-
ford. Cook (2010) notes that this flow pathway is approximately sub-parallel to the Hoddesdon Syncline
which points to a structural influence on the flow regime.
A number of public supply abstraction wells show connections to the karst network. These are
generally located relatively close to known springs (i.e. Broadmeads PWS and Chadwell Spring, Amwell
3.2. Geology and Hydrogeology of Hertfordshire 76
Marsh PWS and Emma’s Well, Rye Common and Rye House Spring, and Hoddesdon/Middlefield Road
and Lynchmill Spring) and the majority have extensive adit systems. It is likely that long-term operation
and aquifer development around the abstraction wells has encouraged the convergence of rapid flow
paths to these discharge points.
The tracing by Cook (2010) also indicated that rapid flow paths appear to extend further west
beyond the zone of main karst development into the Vale of St. Albans. Surface karst features are
not apparent within the Vale of St. Albans due to the extensive glacial deposits covering the Chalk.
The karst system in the Vale of St. Albans appears to be less continuous and less well developed than
the main karst network along the Palaeocene feather edge, perhaps restricted as a result of infilling,
weathering and/or erosion of karst features with increasing distance from the Palaeocene feather edge
(or within ‘Geomorphic Zone 2’ of Maurice et al. (2006)). Flow velocities in this area are more variable,
and generally lower. The breakthroughs suggest a general decline in velocity and mass recovery with
distance from the Palaeocene outcrop. Transport is significantly attenuated, perhaps suggesting multiple
flow paths reflecting a range of dissolution enhanced fissure sizes. This higher attenuation may also, at
least partly, be a consequence of the tracer injection locations being boreholes, and therefore not directly
connected to the conduit system, resulting in dilution and dispersion before the karst system is reached
(Worthington, 2003; White, 2003). In addition to the tracer evidence, small conduits have been observed
in boreholes in the northern part of St. Albans CL:AIRE (2002).
Tracer was not detected at Chadwell Spring in the recent tracer tests (Cook, 2010), despite being
detected in the 1920s and 1930s tests. Cook (2010) suggests that flow to this spring has been affected by
recent changes in abstraction patterns and it now derives water from further north, outside the catchment
area of the tracer. This is supported by evidence from water chemistry (Section 3.4). This could also
explain why bromate concentrations are typically lower at Chadwell Spring than at other locations to the
south along the Lea Valley.
Cook (2010) proposes that the pattern of tracer breakthroughs could suggest that the Northern Loop
of the Palaeocene Feather Edge and River Lee is short-cut by the subsurface karst system, with a more
direct karst component to the central Lea Valley, and a dispersive component beyond Arkley Hole to the
northern Lea Valley. This could provide the mechanism by which the bromate concentrations tend to
be higher in the more southerly Wells than the NNR Wells since they receive a more direct and a less
diluted karst component.
In the area between North Mimms and Tyttenhanger Park, and the Ellenbrook, the Chalk-PTG
groundwater aquifer is separated from the river by a continuous (although <3 m in places) layer of
Boulder Clay. The shallow perched sand and gravel aquifer overlying the boulder clay is considered to
be in hydraulic continuity with the surface water system. This system is likely to have been influenced
by gravel extraction activities, which may have locally increased the connectivity between the Chalk
aquifer and the river.
Boulder Clay is absent in some areas over the section downstream of Tyttenhanger Park to the
confluence with the River Ver, which suggests potential for greater connection between the river and the
Chalk-PTG groundwater system. The river may be influent during periods of high groundwater levels
and effluent over some sections during periods of low water levels.
3.3 Piezometry
Long-term water level data is available for a number of observation boreholes across the catchment
(Figure 3.5). These boreholes form part of the Environment Agency monitoring network. The long-term
rainfall and water level variations are shown in Figure 3.6, and a detailed time series for Orchard Garage
monitoring well, located close to the source site in Sandridge is shown in Figure 3.7.
Details of the Chalk piezometry, groundwater movement and water level fluctuations in the area
have been collated and described by Entec (2000), Buckle (2002) and Atkins (2004). In general, water
levels in the Chalk respond to variations in rainfall. Locations in the western part of the study area, on
the upland interfluve between the River Lea and River Ver show the largest fluctuations (approximately
5 m to 8 m), and locations nearer to the river valleys (e.g. in the west of the study area near the River
Colne, or in the east of the study area, close to the River Lea and River Mimram) show less pronounced
±
TL11/6
_
^
TL31/1
_
^
TL31/98A
TL11/55 TL31/3 _
^
_
^ TL20/6D _
^
_
^
TL10/50 TL10/8A
_
^ TL10/14 _
^
_
^
TL20/14 TL20/49
_
^ _
^
TL10/63
_
^
3.3. Piezometry
_
^ Environment Agency long-term observation boreholes
Figure 3.5: Environment Agency monitoring network long-term water level monitoring locations
200
100
160
120
80
40
0
Total Monthly Rainfall (mm)
(Rothamsted Station)
TL 11/6
80
TL 10/50
TL 11/55
TL 10/14
TL 20/49
TL 10/8A
TL 10/63
TL 19/19C
60
TL 20/14
TL_11_6
TL_10_50
TL_11_55
TL 20/6D TL_10_14
TL_10_8A
TL 31/1 TL_10_63
40
TL_20_49
TL 31/3 TL_19_19C
TL_20_14
TL 31/98A TL_20_6D
TL_31_1
TL_31_3
TL_31_98A
20
Figure 3.6: Rainfall and Environment Agency monitoring network water level variations. Locations of monitoring wells are shown in Figure 3.5
3.3. Piezometry 81
fluctuations (typically less than 3 m). The presence of sand and gravel drift deposits above the chalk is
also likely to reduce water level fluctuations as a result of greater storage capacity.
Notable fluctuations in water levels are summarised below:
• Between 1990 and 1992, water levels fell to some of their lowest recorded levels, following a
period of relatively high water levels between 1981 and 1988.
• Between 2000 and 2003 (particularly 2001), water levels rose to the maximum recorded levels.
• Between 2003 and 2006, water levels declined but started to rise again over 2007.
200
Orchard Garage TL 11/55
SMD
80
40
82
80
78
76
74
Figure 3.7: Water Level, Soil Moisture Deficit and rainfall at Orchard Garage Monitoring Well
A series of piezometric contour maps have been produced by Entec (2000), Buckle (2002) and
Atkins (2004). The maps show that the overall pattern of piezometry has remained reasonably constant
over the 1990s and early 2000s (although data up until 1997 is less extensive than later dates, particularly
in the areas between Sandridge and Hatfield). There is some uncertainty in delineating contour lines
where insufficient data points were available, notably between Tyttenhanger PS and Netherwild PS. It
is thought likely that the most significant water balance shifts in the Upper Lea area probably occurred
prior to 1970 as a result of significant changes in abstraction regimes. The long-term average piezometry
is shown in Figure 3.8.
On the basis of the piezometric data, groundwater flow tends to follow topography away from
the Chalk upland in the north and north-west, where groundwater elevations can exceed +100 m OD,
towards the confined Chalk in the south and south-east, where groundwater elevations are less than
+20 m OD. Within the Lee Valley in the vicinity of the NNR sources, flow directions tend to follow the
course of the river such that groundwater generally flows from north to south (Atkins, 2004). Ground-
water elevations continue to fall south-east towards the Lee Valley and ultimately south towards Central
Broadmeads P.S.
!
A 80 !A!
± 78 !A
A
76
74 70 66 Chadwell Spring
72 A!
68 64 !
Amwell Hill P.S. A
62
60 !
A
58 56 Essendon P.S. Rye Common P.S.
50
! Waterhall P.S. Middlefield Road P.S.
A !
A !
A
A!
Stonecross P.S. Bishops Rise PS.
Hatfield P.S. Hoddesdon P.S.
!
A ! 45
Holywell PS Raw
A !
A
Broxbourne P.S.
! Roestock P.S. 40 30
A Tyttenhanger P.S.
56
!
! A
A
Turnford P.S.
3.3. Piezometry
Figure 3.8: Average piezometry 1998 to 2008. Contour levels are in m AOD. Arrows indicate groundwater flow direction
3.3. Piezometry 83
London. There is a groundwater divide between units contributing to the River Lee and those contribut-
ing to the River Colne. This divide runs through the central part of the study area between Sandridge
and North Mymms.
The Radlett–North Mymms area recharge mound can be seen on piezometric maps as a groundwater
high aligned along the Palaeogene escarpment. The London Clay cover is absent in places, exposing the
Chalk and the sands and clays of the Reading Formation, thus allowing recharge to the Chalk aquifer to
occur more readily in this area compared to the surrounding areas. Recharge is also contributed to by
the presence of swallow holes in this area, notably the Water End complex. This feature is seen to be
present even during extended periods of low rainfall. Groundwater moves in all directions away from
the mound, although the greatest flow is towards the north-east. Piezometry and groundwater flow in this
area is not well defined, and is considered to be sensitive to changes in recharge Buckle (2002). Water
flow in this area is also influenced by the abstractions to the northwest and west of the recharge mound.
The swallow holes in the North Mymms-Water End area are discussed further in Section 3.2.5.
Walsh and Ockenden (1982) note that the there is an apparent barrier to flow coincident with the Shenley-
Broxbourne anticline, and a preferential flow direction along the line of the Colney Heath-Hoddesdon
syncline. This may indicate a high permeability conduit or zone linking the mound and the River Lee.
Buckle (2002) invokes the presence of such a zone, straddling the zone of water table fluctuation and
underlain by a lower permeability main body of Chalk, to explain the persistence of the mound. When
operative under high flow conditions, the zone allows removal of the groundwater from the recharge area.
However, under low recharge conditions, the recharge mound is maintained by the lower transmissivity
of the main body of Chalk in which flows are much reduced.
Sandridge is situated at the confluence of two dry valleys, and the St Leonard’s Court (SLC) source site
itself is situated on the northern edge of the valley down from the confluence. The dry valley trends
south-east from Sandridge, following the general trend of a major joint/fracture set. Roberts (2001)
suggests that preferential flow paths may have developed within the upper levels of the Chalk in the
Sandridge area as the dry valley represents a local discharge area for the Chalk aquifer, and it was
probably also a tributary of the former course of the Thames through the Vale of St Albans. Piezometric
maps show that in the Sandridge area groundwater movement is in general from west-north-west to east-
south-east. However, information is not available in sufficient detail to allow more precise delineation
of flow direction along the line of the dry valley. The groundwater flow in the vicinity of the SLC site is
discussed further in the Chapter 5.
3.4. Regional hydrochemisty 84
Type IA Background regional waters, characterised by calcium bicarbonate waters with low potassium
and magnesium concentrations. Occur extensively to the north of the River Lee and to the north
and west of Ware.
Type II Recharge/surface waters, with low alkalinity and high pH, nitrate, suplphate, chloride and
sodium concentrations. Grouped around the North Mimms area swallow holes and represent rapid
recharge waters.
Type III Mature groundwaters. Typically present in the confined section of the aquifer and have char-
acteristically very low to negligible nitrate. Found in the confined section of the Chalk.
Type V Mixed groundwaters. These have characteristics intermediate between Type IA and Type II
waters. Wide spatial scatter and occur throughout the region, but found mainly at locations south
and west of the River Lee in the areas known to be affected by rapid recharge mechanisms.
The groundwater abstracted from the NNR sources appears to contain varying proportions of Type
IA and Type II groundwaters. Furthermore, the Type IA groundwaters can be further distinguished as
those derived from the north of the River Lee towards Ware and Stevenage (Type IA-N), and those de-
rived from the west of the area around Sandridge (Type IA-W). Atkins (2004) point out that determining
3.5. Scavenge Pumping at Hatfield Pumping Station 85
the proportion of Type 1A-W water (which contains bromate) contributing to each of the sources would
be an effective way of estimating likely bromate concentrations. To do this would require that Type
IA-N waters could be more readily distinguished from Type IA-W waters, for example on the basis of a
contaminant (other than bromate) present in one body of water but not the other.
The data were interpreted as suggesting that there is seasonal variation in dominant flow directions
between the Water End swallow holes, the Essendon area and the Lee valley in the vicinity of the NNR
wellfield. During the winter months flow directions away from the Essendon area appear to be predomi-
nantly northeast or east, whilst in the summer months they have less of a north-easterly component.
• to assess whether there is a statistically significant relationship between Hatfield abstraction rate
and bromate concentration at Essendon and the NNR sources;
The Hatfield source has been pumped at rates up to 9 Ml day−1 while bromate, bromide, sulphate
and chloride concentrations have been monitored at the TVW Hatfield and Essendon sources and the
TWUL Northern New River Sources (see Figure 3.1 for locations). Pumping has continued into 2009.
Initial analysis as an internal report to TWUL and TVW (Fitzpatrick, 2007) considered data up until the
end of December 2006 only. This thesis considers the data up until the end of December 2008.
3.5. Scavenge Pumping at Hatfield Pumping Station 86
160
80
0
40
Hatfied Abstraction Rate (litres per day)
6
9.0x10
8.0x106 0
6
7.0x10
6
6.0x10
5.0x106
6
4.0x10
3.0x106
2.0x106
1.0x106
0
0.0x10
Figure 3.9: Abstraction rates at Hatfield PS between 31 June 2005 and 31 December 2008
following months. A ‘deseasonalised’ trend was estimated separately for data before and after the start
of the pumping trial. This was estimated following the method outlined in Chatfield (2004.):
1
2 x(t−6) + x(t−5) + . . . + x(t+5) + 21 x(t+6)
Sm (xt ) = (3.1)
12
where xt is the average monthly concentration at time t and Sm (xt ) is the deseasonalised component of
xt .
200
160 160
120
120
SMD
80
80
40
40
0
0
8
400
4
200
0 0
Raw Br 10
Concentration (µg l-1)
8
800
4
400
0 0
Figure 3.10: Time series of bromate and bromide concentrations at Hatfield PS, soil moisture deficit, and
monthly rainfall.
3.5. Scavenge Pumping at Hatfield Pumping Station 89
200
160 160
120 120
SMD
80
80
40
40
0
0
40
8
20
4
0 0
Raw Br 10
Concentration (µg l-1)
200
8
100
4
0 0
Figure 3.11: Time series of bromate and bromide concentrations at Essendon PS, soil moisture deficit,
and monthly rainfall.
3.5. Scavenge Pumping at Hatfield Pumping Station 90
200
Monthly Rainfall (mm)
Rothamsted Station
160
SMD
80
80
40
40
0
0
6 8
6
4
2
2
0 0
Raw Br 10
Concentration (µg l-1)
200
8
100
4
0 0
Figure 3.12: Time series of bromate and bromide concentrations at Chadwell Spring, soil moisture
deficit, and monthly rainfall.
3.5. Scavenge Pumping at Hatfield Pumping Station 91
200
Monthly Rainfall (mm)
Rothamsted Station
160
SMD
80
80
40
40
0
0
30 8
6
20
10
2
0 0
Raw Br 10
Concentration (µg l-1)
200
8
100
4
0 0
Figure 3.13: Time series of bromate and bromide concentrations at Amwell Hill PS, soil moisture deficit,
and monthly rainfall.
3.5. Scavenge Pumping at Hatfield Pumping Station 92
200
Monthly Rainfall (mm)
Rothamsted Station
160
SMD
80
80
40
40
0
0
30 8
6
20
10
2
0 0
Raw Br 10
Concentration (µg l-1)
200
8
100
4
0 0
Figure 3.14: Time series of bromate and bromide concentrations at Amwell Marsh PS, soil moisture
deficit, and monthly rainfall.
3.5. Scavenge Pumping at Hatfield Pumping Station 93
200
Monthly Rainfall (mm)
Rothamsted Station
160
SMD
80
80
40
40
0
0
20
8
10
4
0 0
Raw Br 10
Concentration (µg l-1)
200
8
100
4
0 0
Figure 3.15: Time series of bromate and bromide concentrations at Rye Common PS, soil moisture
deficit, and monthly rainfall.
3.5. Scavenge Pumping at Hatfield Pumping Station 94
200
Monthly Rainfall (mm)
Rothamsted Station
160
SMD
80
80
40
40
0
0
30 8
6
20
10
2
0 0
Raw Br 10
Concentration (µg l-1)
200
8
100
4
0 0
Figure 3.16: Time series of bromate and bromide concentrations at Middlefield Road PS, soil moisture
deficit, and monthly rainfall.
3.5. Scavenge Pumping at Hatfield Pumping Station 95
200
Monthly Rainfall (mm)
Rothamsted Station
160
SMD
80
80
40
40
0
0
60 8
6
40
20
2
0 0
300 Raw Br 10
Concentration (µg l-1)
200
6
4
100
0 0
Figure 3.17: Time series of bromate and bromide concentrations at Hoddesdon PS, soil moisture deficit,
and monthly rainfall.
3.5. Scavenge Pumping at Hatfield Pumping Station 96
200
Monthly Rainfall (mm)
Rothamsted Station
160
SMD
80
80
40
40
0
0
60 8
6
40
20
2
0 0
300 Raw Br 10
Concentration (µg l-1)
200
6
4
100
0 0
Figure 3.18: Time series of bromate and bromide concentrations at Broxbourne PS, soil moisture deficit,
and monthly rainfall.
3.5. Scavenge Pumping at Hatfield Pumping Station 97
200
Monthly Rainfall (mm)
Rothamsted Station
SMD
80
80
40
40
0
0
40
8
20
4
0 0
300 Raw Br 10
Concentration (µg l-1)
200
6
4
100
0 0
Figure 3.19: Time series of bromate and bromide concentrations at Turnford PS, soil moisture deficit,
and monthly rainfall.
3.5. Scavenge Pumping at Hatfield Pumping Station 98
ure 3.13), Amwell Marsh (Figure 3.14) and Rye Common (Figure 3.15) sources follow a similar pattern.
Maximum concentrations for each cycle occur in the spring months (March, April, May) and lowest
concentrations occur in the autumn months between September and December. However, a second peak
occurs in 2004 during September, October and November followed by a trough in December or January.
The Hatfield pumping trial begins at the end of July when bromate concentrations are on the downward
limb of the seasonal cycle. The seasonal cycle appears to continue, with peak concentrations occurring
in April and concentrations declining to a trough in October, and appears to continue through 2007 and
2008.
The Hoddesdon (Figure 3.17), Broxbourne (Figure 3.18) and Turnford (Figure 3.19) sources show
a slightly different seasonal trend. Maximum concentrations occur slightly later in summer and early
autumn (June, July, August, September) with a more rapid decline to minimum concentrations in winter
(November, December, January). The Hatfield pumping trial begins at the end of July when bromate
concentrations are at their peaks. Concentrations appear to be rapidly decreased during pumping. The
seasonal cycle appears to be perturbed, with peaks occurring in April, earlier than expected and during
the period that pumping at Hatfield was suspended, and the falling limb occurring earlier over May and
June when pumping was resumed. However, the seasonal cycle is resumed during the prolonged periods
of pumping in 2007 and 2008.
For Hatfield the deseasonalised trend indicates an overall increasing trend in bromate concentrations
between 2000 and July 2005, although mean concentrations appear to level off in 2006, and there is
an indication of an overall decrease in concentrations by December 2006. Since then the variation in
bromate concentrations about the average trend has increased dramatically. For Essendon, the seasonal
variation is superimposed on an overall rise in bromate concentration, which is clearly indicated by the
‘deseasonalised’ trend line. Between July 2005 and October 2007 there is a general decline in bromate
concentrations with a few peaks, corresponding to periods of non-pumping from Hatfield. Between
October 2007 and December 2008 concentrations appear to level off, with an indication of a slight rise
in bromate concentrations from January 2008.
For the NNR wells the overall trends in bromate concentrations since the beginning of 2002 indicate
a general increase from 2002 until mid 2003. Subsequently, concentrations appear to level off, albeit with
fluctuations. The trend lines then show an increasing trend through the late 2004 and into 2005 (with
the exception of Amwell Marsh where concentrations continue to decline into 2005). Concentrations
generally fall with the start of pumping from Hatfield at the end of July 2006. In February and March
2006 all NNR sources show a significant rise corresponding to the period when pumping was suspended
at Hatfield. Concentrations fall once again in May 2006 following the Hatfield switch on. The trend lines
for bromate concentrations show a general increase from late 2006-early 2007 until the end of 2008, with
concentrations approaching those seen prior to the start of the pumping trial (and higher in the case of
Middlefield Road).
3.5. Scavenge Pumping at Hatfield Pumping Station 99
Table 3.5: Pearson correlation coefficients for Bromate concentration and Soil Moisture Deficit (SMD)
before the start of the Hatfield pumping trial on 29th July 2005. SMD-X corresponds to the SMD value
X months previously. Shaded cells indicate the strongest relationship.
After the start of pumping at Hatfield, for the majority of the monitored source wells, the relation-
ship between bromate concentration and SMD is not as strong, and in many cases appears contrary to
previous years. This is especially noticeable at Essendon, Hoddesdon, Broxbourne and Turnford where
a peak in bromate concentration between March and May 2006 is associated with low SMDs. However,
as discussed further in Section 3.5.6.5, this is also associated with the period when Hatfield was switched
off and therefore likely to indicate a rebound in bromate concentrations which were lowered by pumping
at Hatfield.
Broxbourne which show concentrations between 10 µg l−1 and 50 µg l−1 . Rye Common, Amwell Hill
and Amwell Marsh show concentrations between 5 µg l−1 and 20 µg l−1 . Amwell End, Broadmeads,
and Chadwell Spring show bromate concentrations less than 0.6 µg l−1 for much of the test. These three
sources are the most northerly of the NNR sources. Bromate concentrations at Amwell End remain less
than 0.6 µg l−1 throughout the period of the test. Bromate concentrations at Broadmeads rise above
0.6 µg l−1 on three occasions throughout the test to 1.1, 1.3 and 1.5 µg l−1 , although these rises occur at
separate times which do not appear to be associated with changes in pumping rates. Bromate concentra-
tions at Chadwell Spring drop below detection limits during the periods of high pumping rates, and rise
to maximum of 3.3 µg l−1 in March and April 2006. For these three sources bromide concentrations do
not appear to show a response to the changes in Hatfield abstraction rate.
As with Essendon, the seasonal relationship between bromate concentration and SMD at Hoddes-
don, Broxbourne and Turnford appears to be perturbed by the effect of abstraction at Hatfield. Peaks in
bromate concentration occur in April, earlier than in previous years and during the period when pump-
ing at Hatfield was suspended, and the falling limb occurs over May and June, earlier than in previous
years and when pumping was resumed. For Amwell Hill, Amwell Marsh and Rye Common, the ex-
pected peak concentration occurs during the spring months (March, April, May) which coincides with
the period when Hatfield was switched off.
The abstraction rate at Hatfield for the day corresponding to the bromate and bromide sample was se-
lected. Bromate and bromide concentrations for each source well were compared to abstraction rates
from Hatfield and also abstraction rates from that particular source. Linear regression was undertaken
using Hatfield abstraction rate as a predictor. Statistical summary tables are included in Appendix C.
For the regression analysis, bromate concentrations for the each of the source wells were first com-
pared to Hatfield abstraction rate for the same day as the sample was taken (day T ), and then sequentially
to abstraction rates for the previous days (day T − 1, T − 2, T − 3 etc.). At each stage the strength of the
relationship was assessed by examining the fitted line scatter plot, the F -statistic and associated P -value
to determine the significance of the relationship, the standard error of the regression as a measure of the
dispersion of the data around the regression line, and the R2 value as a measure of the proportion of
the variation in the response (y) variable that is accounted for the variation in the explanatory variable
(x). The combination of these parameters was used to select the time lag for Hatfield pumping rate that
produced the best predictor variable for bromate concentration at each source well. The residuals for this
‘best fit’ regression relationship were examined to assess whether the assumptions of linear regression
(Section 3.5.6.3) were adhered to and therefore if the regression relationship could be used for further
hypothesis testing.
Where the regression relationship was deemed to conform to the assumptions, the hypothesis that
the coefficient of the regression line was significantly different from zero was tested by examining the
t-statistic and associated P -value from the Minitab output.
3.5. Scavenge Pumping at Hatfield Pumping Station 102
• the variance of the residuals is constant (homoscedastic), i.e. it does not depend on the x or on
anything else (e.g. on time);
Plots of residuals versus predicted values can indicate heteroscedasticity of the residuals, and the
normality of residuals can be checked by a normal probability plot. A plot of residuals against time can
indicate correlations between residuals over time.
1. variation in the residuals is larger for larger fitted values (lower pumping rates);
Residuals Adherence to
1
Source Time Lag Linearity of relationship 2 3 4 Assumptions of
Homoscedasticity Independence Normality
Linear Regression?
For equation Bromate = A + B(Hat abst)
Hatfield
Although there is a lot of scatter about the The variation of the
Seasonality is not Residuals follow the
0 days line, a linear relationship describes the residuals is reasonably
evident in the residuals. normal line well.
‘average’ data trend reasonably well. constant.
Including SMD
Figure 3.20: Assessment of residuals for each ’best-fit’ regression for the response of bromate concentration to Hatfield abstraction rate. (1)
Residuals Adherence to
1
Source Time Lag Linearity of relationship 2 3 4 Assumptions of
Homoscedasticity Independence Normality
Linear Regression?
For equation Log(Bromate) = A + B(Hat abst)
Seasonality is evident
with positive residuals in
Chadwell Spring Residuals follow the
Variation is less for smaller the winter/spring and May need to add
There is more variation at low pumping normal line well.
fitted values and less negative residuals in the seasonal term
6 days rates, however the position of the line looks
variation between July summer/autumn.
reasonable.
2005 and June 2006. Residuals are still Residuals show slight
Including SMD associated with date. departure from
Still correlation with date
normality.
Seasonality is evident
The variation of the
with positive residuals in Residuals show
residuals is reasonably
Amwell Hill the winter/spring and departure from May need to add
There is more variation at low pumping constant, although more
negative residuals in the normality. seasonal term
5 days rates, however the position of the line looks for positive residuals.
summer/autumn.
reasonable.
The variation of the Residuals are still
Residuals follow the
Including SMD residuals is reasonably associated with date.
normal line well. Still correlation with date
constant.
Residuals show slight
The variation of the Slight seasonality Reasonable, but may
Middlefield Road departure from
Although there is a lot of scatter about the residuals is reasonably present. need to add seasonal
normality.
8 days line, a linear relationship describes the constant, although slightly term to equation
‘average’ data trend reasonably well. less for smaller fitted Residuals are still
Including SMD values. associated with date. Reasonable, but still
dependent on date.
1
Inspection of fitted line scatter plots
2
Inspection of residuals versus fits plots
3
Inspection of residuals versus order plot
4
Inspection of normal probability plot
3.5. Scavenge Pumping at Hatfield Pumping Station
104
Figure 3.21: Assessment of residuals for each ’best-fit’ regression for the response of bromate concentration to Hatfield abstraction rate. (2)
3.5. Scavenge Pumping at Hatfield Pumping Station 105
Table 3.6: Summary of regression parameters for the ‘best-fit’ regressions for the response of bromate
concentration to Hatfield abstraction rate.
2 2
Time 2 b R R
Source a R Standard error P-value
lag Hat abst SMD
Heteroscedasticity in the residuals was improved in the majority of cases by transforming the data
by taking logarithms of bromate concentrations (as recommended in Helsel and Hirsch 1993).
Seasonality was apparent in the majority of the NNR wells, and at Essendon. Therefore, SMD was
included as a predictor in the regression relationship. In general, including SMD as a predictor in the
regression increased the amount of variation explained by the regression (increased the value of R2 ).
The residuals showed correlation to SMD. Positive residuals, indicating that observed bromate con-
centrations are above the fitted regression line, generally occur when SMD is low (winter and spring),
and negative residuals, indicating that observed bromate concentrations are below the fitted regression
line, generally occur when SMD is high (summer and autumn). This appears to be contrary to the rela-
tionship observed between bromate concentration and SMD prior to the start of the Hatfield pumping test
(Section 3.5.5.2). Positive residuals occur earlier than the expected seasonal peak in bromate concentra-
tions. For Essendon, the normal relationship of positive residuals with low SMD and negative residuals
with high SMD occurred.
However, the residuals are also dependent on date, even after the inclusion of SMD. It is apparent
that the most positive residuals occur mainly during the period between January 2005 and May 2006 and
the period over November 2006 when the Hatfield pumping test was suspended for prolonged periods
and only sporadic abstraction at rates less than 3 Ml day−1 occurred for sampling purposes. The positive
residuals may therefore reflect a rebound in bromate concentrations after a reduction due to the effects
3.5. Scavenge Pumping at Hatfield Pumping Station 106
of pumping. The fact that the positive residuals become less positive and more negative when pumping
was resumed in May 2006, and are negative between August 2005 and December 2005 and between
May 2006 and December 2006 when prolonged periods of higher abstraction rates occurred, indicates
that the effect of Hatfield abstraction rate on bromate concentration appears to dominate the seasonal
relationship between SMD and bromate concentration. The apparently contrary relationship between
bromate concentrations and SMD may therefore be an artefact of the timing of the abstraction rate
variations at Hatfield.
The relative effect of SMD and Hatfield abstraction rate is difficult to measure due to the timing
of the abstraction rate variations at Hatfield. The SMD curves reach their seasonal trough between
January and April 2006. The Hatfield pumping test was suspended between January 2005 and May
2006 and only sporadic abstraction at rates less than 3 Ml day−1 occurred for sampling purposes. The
higher abstraction rates occurred from August 2005 to December 2005 and May 2006 to December 2006
when SMD was in the higher part of its seasonal cycle. This results in an apparent positive correlation
between Hatfield abstraction rate and SMD. In order to separate these two effects it would be necessary
to maintain more constant abstraction rates at Hatfield over the full cycle of SMD variations.
A more constant period of abstraction occurred between November 2007 and June 2008, when ab-
stratction rates were maintained at relatively consistent high values of ∼ 6 Ml day−1 to ∼ 8 Ml day−1 .
Residuals still tend to show seasonal variations, which suggests that other seasonal variables are impor-
tant in controlling bromate concentrations.
The effect of rainfall is also complicated as a result of its relationship to abstraction rate at Hatfield.
Sewer surcharging events in response to heavy rainfall caused Hatfield abstraction to cease. As a result,
the abstraction rate at Hatfield is heavily influenced by rainfall in the catchment.
and Hoddessdon only. Therefore a significant linear correlation exists between bromate concentration
and SMD for these source wells.
Table 3.7: Coefficients determined by the ‘best-fit’ regressions for the response of bromate concentration
to Hatfield abstraction rate.
Figure 3.22 compares the mean and 95% confidence intervals for the coefficients determined by the
linear regression for Essendon, Broxbourne and Hoddesdon. The confidence interval for Essendon is
smaller than for Broxbourne and Hoddesdon and Turnford. The confidence intervals all overlap, indi-
cating that the differences in mean coefficients are not statistically significant at a 95% confidence level.
Therefore, the gradient of the relationships between bromate concentration and Hatfield abstraction rate
for the monitored sources apart from Hatfield do not appear to differ significantly between each source.
0
3.5. Scavenge Pumping at Hatfield Pumping Station
109
Figure 3.23: Comparison of statistical response times for bromate concentration response to hatfield abstraction and tracer travel times from Water End. Based on Cook
(2010)
3.6. Single borehole dilution testing 110
• To determine the hydraulically active horizons within the selected boreholes in order to guide
injection strategies for the natural gradient point-to-point tracer testing;
• To use uniform injection SBDTs to obtain a direct measurement of horizontal specific discharge
(Darcy velocity);
Detailed methodology and interpretation of the results is included in Appendix D. The determina-
tion of horizontal darcy velocities is described within this Section.
Horizontal specific discharge (Darcy Velocity) was determined (according the the methodology
outlined in Figure 3.24) for three locations within the study area (location numbers refer to Figure 4.1):
Figure 3.24: Methodology for determination of specific discharge (darcy velocity) from the results of
the Single Borehole Dilution Tests. Based on Ward et al. (1998)
3.6. Single borehole dilution testing 112
-20.50
-21.50
-22.50
-23.50
-24.50
Depth (mbD)
-25.50
-26.50
-27.50
-28.50
-29.50
-30.50
Figure 3.25: Specific discharge (darcy velocity) for each 0.5 m depth section at Nashes Farm. Estimated
Ct −Cb
using the methodology in Figure 3.24. Plots of ln C0 −Cb
are included in Appendix D. The value at each
section is estimated based on Based on Single Borehole Dilution test carried out at Nashes Farm 29
January 2008.
3.7. Conceptual Model for groundwater flow in Hertfordshire 113
-16.0
-16.4
-16.8
-17.2
-17.6
Depth (mbD)
-18.0
-18.4
-18.8
-19.2
Figure 3.26: Specific discharge (darcy velocity) for each 0.5 m depth section at Comet Way BH. Esti-
Ct −Cb
mated using the methodology in Figure 3.24. Plots of ln C 0 −Cb
are included in Appendix D. The value
at each section is estimated based on Based on Single Borehole Dilution test carried out at Comet Way
BH 4 February 2008.
-16.0
-17.0
-18.0
-19.0
-20.0
Depth (mbD)
-21.0
-22.0
-23.0
-24.0
-25.0
-26.0
Figure 3.27: Specific discharge (darcy velocity) for each 0.5 m depth section at Harefield House BH.
Ct −Cb
Estimated using the methodology in Figure 3.24. Plots of ln C0 −Cb
are included in Appendix D. The
value at each section is estimated based on Based on Single Borehole Dilution test carried out at Harefield
House BH on 22 January 2008.
wellfield. Transport in the karst conduits is characterised by low attenuation and high flow rates. There
is also likely to be a background flow of bromate contaminated water east of Hatfield that is influenced
by double-porosity characteristics.
Figure 3.28: Conceptual model for groundwater flow in the bromate affected area of Hertfordshire. Position of conduits are based on the conceptual model developed by
Cook (2010). Flow rates and attenuation characteristics are inferred from the results of the single borehole dilution testing presented in Section 3.6 and tracer tests undertaken
by Cook (2010).
3.8. Summary and conclusions 116
the Hatfield area and springs and abstraction wells in the Lea Valley. In combination with tracer testing
undertaken by Cook (2010), this suggests a conduit system which connects three regions Water End,
Hatfield, and the northern and middle Lea Valley. The results of the tracer testing from Water End
indicate rapid flows of the order of 1000 m day−1 along the Palaeocene boundary.
The information assessed and interpreted in this chapter has been used to develop a conceptual
model for flow and transport within the Hertfordshire Chalk aquifer which considers that double-porosity
characteristics dominate close to the source site in Sandridge and within the Vale of St Albans, resulting
in high attenuation of bromate. The main karst network is developed along the Palaeocene boundary and
allows rapid transport of bromate, with low attenutation, toward the Lea Valley.
117
Chapter 4
sheet. (Initially the monitoring data was held by a number of different organisations undertaking separate
monitoring.) The Environment Agency assumed the role of data management and transfer of the data
occurred in September/October 2001. The data was entered in the Agency’s WIMS database, and from
this the data was subsequently exported into an Access database (the ‘Bromate Monitoring Database’
). Although at the time, the majority of samples taken were being analysed at the VWP laboratory it
was always the intention to eventually transfer sample analysis to the Agency laboratory (this occurred
in October 2002). Automatic transfer of data occurs from the laboratory to the WIMS database. The
EA continues to manage all publically available monitoring data (some public water supply data are
excluded) within the database, including both water quality results and water levels. The database also
includes a variety of other information including monitoring location reference information (e.g. NGR,
owner, borehole depth, topographic elevation etc) and sample schedules (‘runs’).
In September 2005, the Environment Agency commissioned Atkins Limited to produce a factual
and interpretative report (Atkins, 2006) of the data and associated information obtained through the five
years of monitoring. In addition, Atkins amended/upgraded the monitoring database. In order to form
the basis of the data presented in this thesis chapter, the Bromate Monitoring Database (including all
data up to the end of December 2008) was provided by the Environment Agency. Monitoring data (up to
the end of December 2008) provided by TVW and TWUL was included within a modified version of the
database, the ‘UCL Bromate Monitoring Database’. The UCL database was linked to a GIS database.
148 A!A
201 !202
A!A 152 A!A ! 005
A
153 001197
! !A
!A
! !
A!205 235 294 315 299
A A
!A
A! ! !
200204 A ! 078
!
A A A A!
312313314 !
A
348 A
349! A! 171195
196 345A!
! 010b ! 170 A A
!316
!!
!
A
A!347 A 057
A
A 346 021102 192 A!
141 193 142 !
A
A!
! A!
!
011A! A 049
012013 !
A 086
!! 015 302
A !
A
!
A 006 !
014 A
! ! 140
A A !
A
145
0 A
! 2 4 6 8 10
Kilometers
119
Table 4.1: Typical analytical suite for water samples May 2000 to December 2008
Table 4.2: Analytical methodology and detection limits for bromate analyses.
For the period May 2000 to late October 2002, all samples were analysed by the VWP laboratory.
During this time, from January 2001, new analytical methods for bromate and bromide were introduced.
A duplicate analysis was undertaken to compare the old and new analytical method. This involved the
collection and analysis of approximately 20 field samples. Initial testing showed that bromate results
were 20 % lower for concentrations above 150 µg l−1 .
According to Buckle (2002), the laboratory uses an ion-chromatographic technique, and has NA-
MAS accreditation for the parameters analysed. The occurrence of ‘isolated and unexpected’ bromate
4.3. Monitoring Data Quality 121
sample results, particularly for TVW operated sources (e.g. Roestock and North Mymms), recorded
above the detection limit of 1 µg l−1 , but generally less than 5 µg l−1 , has shown that the technique
is not completely reliable at low concentrations. Under normal circumstances, bromate and bromide
concentrations over the detection limit are subject to an estimated measurement error of ±10 %.
From October 2002, all samples were collected by the EA were analysed by the Environment
Agency’s Starcross Laboratory in Exeter. During the period of changeover, duplicates (comprising some
20 field samples) were analysed by both the Starcross laboratory and VWP to check for consistency.
According to the EA, the results of these comparisons indicated that discrepancies resulting from dif-
ferent methods or laboratories were unlikely to exceed 20 % for bromate and bromide concentrations
greater than about 5 µg l−1 . (It should be noted that the comparisons were not intended to be rigorous
statistical exercises, but to provide reassurance that discrepancies resulting from different laboratories
or methods would not significantly distort the broad picture of the distribution of bromate and bromide,
whose concentrations ranged over there orders of magnitude within the plume.)
Further to the above, in June 2005 an inter-laboratory comparison exercise was undertaken to check
the consistency of results. Triplicate samples were analysed by VWP, TWUL and EA laboratories.
Eleven samples were taken for the NNR wells and from TVW sources at Hatfield, Bishops Rise and
Essendon. A comparison exercise was undertaken for the three laboratories, similar to that described
above. Over the concentration range 10 µg l−1 to 350 µg l−1 the maximum deviation of an individual
sample from the mean of three samples was 23 %. For one sample close to the MDL results varied from
<0.5 µg l−1 to 2.3 µg l−1 . Results for bromide showed less variation.
• Phase 1 – An initial phase with the primary objective of identifying the source of contamination
and the extent of the affected area. This phase was effectively achieved by the end of 2000;
• Phase 2 – A subsequent on-going phase to monitor concentrations and assess boundary migration.
!
A
!
A
!A! F
G
A !
!! !
A !
A A A ! A
A!A! !
A
A !
A
! !
A
!
A !
A
!
!
A
A ! !!
A A !
!
A
!! !! !A
A
!A!
A !AA! !
A F
G F
! ! G
A AAA
F
G A!A !
A F
G !
Source Site !
A A A
! A GG
F F !GA! !
A
A !
AF
!
A FA
!G
A
!
A
!
A
! !
A
G
F
A
F
G
!
F
G
!
A A
!
A
Bromate G
F !
F
G A
Monitoring !
A !
A F
G
Frequency
! F
G F
G
A PWS_sym G
F
F
G F
G
G
F !
A
F
G
Yr_2000 !
A
!
A F
G !
A
!
A
!
A 1 - 20
F
G !
A
!
A 21 - 75 !
!
A
A !
A
!
A
!
A 76 - 180
!
A !
A
Groundwater_sym !
A
4.3. Monitoring Data Quality
Yr_2000 F
G
!
A 1-4 !
A
! 5 - 10
A
!
!
A
A
! 11 - 27
A
!
A
Surface water !
A
Yr_2000 !
A
F
G 1-5
Boundary of Source Site
F
G 6 - 10 !
A
Former Works
F 11 - 25
G Building Footprint
0 1 2 4 6 8 10
Kilometers
122
Figure 4.2: Sampling frequency for bromate at each monitoring location in 2000
F
G !
A
!
A
F
G
!
A
± !
A
!
A
!
A F
G
!
A
!
A
!
A
!
A
! !
!
!
!
!
A!
!
A ! AA
!
A
A
A
! ! !!
A A
Source Site
! AA
A! ! A
A
A
!!!A
!!
A ! ! !
A
!
G
FA
!A A !A
A
!! A
! A A F
G G
F
!G
AF !
A
! !
!GF F
G
!
A
A!AA
F
G !
A A
!
A
!F
GA A! !
A !
A
!
A! A
!
G
F
A! A
A ! !
A
F
G
! A!
A F
! G
A !
F
G !
! A
A F
G A
A
! !A
A F
G
!
!A
!
A !A G
Bromate A
! A
A! F
G !
A
!
F A!
! A !
A
Monitoring A
G !
Frequency A! FA
!
A !
A
F
G !
A
!
A PWS F
G
G G !
A
Yr_2001 ! !
A
A
FF
! G
A
F AA!!
F
G
A!
!
!
A 1 - 20
F
G !
A
! 21 - 75
A
!
A
!
A
A! 76 - 180
!
A
Groundwater !
A
4.3. Monitoring Data Quality
Yr_2001 !
A
!
A 1-4 !
! A !
A
!
A !
A !
A !
A 5 - 10 A
!
A
!
A
!! 11 - 27 !
A!
A
AA
!
A
Surface water
!
A !
A
Yr_2001 !
A
F
G 1-5
Boundary of Source Site
F
G 6 - 10
Former Works !
A
Building Footprint
F 11 - 25
G
0 1 2 4 6 8 10
Kilometers
123
Figure 4.3: Sampling frequency for bromate at each monitoring location in 2001
!
A
G
F
!
A
F
G
± !
A
!
A
F
G
F
G
!
A
A! A! !
A
A!
!
A G
FG
FG F
G
F
! A ! !! F
G G
F
!
!A
A ! AA! A
! A ! F
G F
GF
G F
G
!
! ! AA
A
Source Site A ! A !
A !
A !G
F !
A G
F
A
!G
F !
F
GA!
!A A
!AG !G
AG
FGF!A !
A
F!!
F
A A
! AF
G A!A
A !
A !
A
!
F
G
A ! ! !
A ! F
G F
A!A
GG
F
A !
! ! GF GFA
AA F
G
! A
F
G A
Bromate !
A ! G
F
A !
A !
A !
A
Monitoring !
Frequency !
A A
! F
G
A F
G
PWS
G
FGF
Yr_2002 ! G F !
A
A FA
G !
!
A 1 - 20
F
G !
! 21 - 75
F
G A
A !
A
!
A
!
A 76 - 180
!
A
Groundwater
4.3. Monitoring Data Quality
Yr_2002
!
A
!
A 1-4
A
! A !
!A
!
A 5 - 10 A! A! !
! !
A!A !
A
A 11 - 27 AA! !
Surface water
!
A !
A
A
Yr_2002 A!
F
G 1-5
Boundary of Source Site
F
G 6 - 10
Former Works !
Building Footprint A
F 11 - 25
G F
G
0 1 2 4 6 8 10
Kilometers
124
Figure 4.4: Sampling frequency for bromate at each monitoring location in 2002
G
F
!
A
AF
G
!
!
A
! !
A
!A
A F
G
!
± !A
A
!
A!
!
A
!
A
F
G A
!
A
!
A
!
A ! !
A A
!
A !
!
A ! A F
G
A !
A
!
!
A ! !
A
A
A !
A ! !
A !
A F
G
A !
A G
F
!
Source Site
A! ! A! !
A !
A F
G
!G
AF !
F
G
A
!
A F
G !
A A
A
!
A !
A
!
A !
A !
! F
G A
A
F
GG
F
Bromate ! !
A A !
Monitoring A !
A
!
A
Frequency !
A
F
G !
A !
!
A A
PWS !
A
Yr_2003 !
A
!
A
!
A 1 - 20
!
A 21 - 75 F
G
! 76 - 180 !
A
A
!
A
Groundwater
4.3. Monitoring Data Quality
Yr_2003
!
A 1-4 !
A
! 5 - 10
A !
A
!
A 11 - 27
!
A
Surface water !
A
Yr_2003
F
G 1-5
Boundary of Source Site
F
G 6 - 10
Former Works
Building Footprint
F 11 - 25
G
0 1 2 4 6 8 10
Kilometers
125
Figure 4.5: Sampling frequency for bromate at each monitoring location in 2003
F
G
F
G
!
A
!
A
± F
G
!
A
!
A !
A
!
A
!
A !
A
A!
!
!
!
AA !
A! !
A !
A
A F
G F
G
F
G F
G
Source Site ! ! ! G
A F !
A !G
F
A!
A !
A A A !
A! A
!
A
!
A
!
A
!
A !
! ! F
G
A
A A ! !
A A
F
G
!A! F
G
A F
G
Bromate !
A !
A !
A
Monitoring !
A
!
A !
A
Frequency !
! F
G
A
!
A
A! !
A! A
PWS
Yr_2004
! 1 - 20
GA
F
!
A
!
A 21 - 75
!
A
! 76 - 180 !
A
A !
A !
A
Groundwater
4.3. Monitoring Data Quality
F
G Yr_2004
!
A 1-4 !
A
! 5 - 10
A !
A
!
A 11 - 27
!
A
Surface water !
A
Yr_2004
F
G 1-5
Boundary of Source Site
F
G 6 - 10
Former Works !
A
Building Footprint
F 11 - 25
G
0 1 2 4 6 8 10
Kilometers
126
Figure 4.6: Sampling frequency for bromate at each monitoring location in 2004
!
! A
A
F
G
!
A
± !
A F
G
!
A
!
A ! !
!
A A A
! !
A !
A !
A
!
A A ! !
A A A !
! A
!
!
!
A
A! A !
! A ! ! !
A
A
!
A
!!! A A
A !
A
! !
A !
AA
! !
AA ! !
!A
! ! !
! AA A A
A
A A F
G !
A F
G
G ! ! ! !
Source Site FA A !
! A A FA
G !G
F
A! A
! A
F !
A ! !
G !A
A A A
A !
!A !
A
!
A ! F
G
A
! !
A !
A
!
A A !
A
! !
A F
G
A !
A
!
A
!
A !
F
G
Bromate A
! !
A
!
A A !
Monitoring !
A A
Frequency F
G
!
A
PWS
Yr_2005
!
A 1 - 20
!
A
!
A 21 - 75 F
G !
A
!
A
! 76 - 180 !
A
A !
A !
A
F
G
Groundwater
4.3. Monitoring Data Quality
Yr_2005
!
A 1-4 !
A !
A
!
A
! 5 - 10
A !
A
!
A
!
A
! 11 - 27
A F
G
!
A
!
A
Surface water !
A
!
A
Yr_2005
F
G 1-5
Boundary of Source Site
F
G 6 - 10
Former Works !
A
Building Footprint
F 11 - 25
G
0 1 2 4 6 8 10
Kilometers
127
Figure 4.7: Sampling frequency for bromate at each monitoring location in 2005
!
! A
A
!
A
!
± A !
A
!
A
F
G
!
A
!
A
!
A
A! ! !
!
A A A
! !
A !
A
!
A A ! !
A A !
A
!
!
A
A!
!
! !
A ! !
A
!
!A
! A
A
AA! F
G !
! !
A !
! !
! ! !
!
A A
A AA A AAA !
A
!
A ! F
G G
F G
F
Source Site G
FA!! !
A !
A !G
AF
A A G F A ! !
!
A !
A A
A
!A!
!
A F
!G
A
!
A !
! !
A
!
A F
G A !
A
! !
A
A G
F !
A
!
A G
FGF
!
A !
F
G F
G
Bromate A
! !
A
Monitoring A
! !
A A !
A
Frequency F
G
!
A
PWS
Yr_2006 !
A
!
A 1 - 20
!
A
!
A 21 - 75 F
G !
A
!
A
! 76 - 180 F
G !
A
A ! !
A
A F
G
Groundwater
4.3. Monitoring Data Quality
Yr_2006
!
A 1-4 !
A
!
A
! 5 - 10
A !
A
! !
A
A 11 - 27 F
G !
A
!
A
Surface water !
A
Yr_2006 !
A
F
G 1-5
Boundary of Source Site
F
G 6 - 10
Former Works !
A
Building Footprint
F 11 - 25
G
0 1 2 4 6 8 10
Kilometers
128
Figure 4.8: Sampling frequency for bromate at each monitoring location in 2006
!
! A
A
!
A
!
± A !
A
! G
A F !
A
!
A
!
A !
A
! !
A !
A !
A !
!
A A ! !
A A
A
!
!
A!
! A !
! A ! ! !
A
A
!
A
AA
!!
A A A
!
A !
! A F
G !
A !
A !A
!
A !
! A! A ! A
!
A A A A! F
G !
A F
G
! !
A
! !
A A ! !G
F
Source Site F ! A
! A !
A A
G !
GAA F !A
A !
A !
!
A
A
!A!
!
A
!
A ! F
G A !
A A
! !
A !
A !
A !
A
A
!
A F
G !
! !
A A
A ! !
A G
F
A F
G
!
A !
F
G F
G
A
Bromate ! !
A
A
! !
A A !
A
Monitoring
Frequency !
F
G
A
Surface water_sym
!
A
Yr_2007
F
G 1-5 !
A
F
G 6 - 10 !
A
!
A
!
A
F 11 - 25
G !
A !
A
Groundwater
4.3. Monitoring Data Quality
Yr_2007
!
A 1-4 !
A
! !
A
A 5 - 10
!
A
!
A
!
A 11 - 27
!
A
!
A
PWS !
A
Yr_2007 !
A
!
A 1 - 20
! Boundary of Source Site
A 21 - 75
Former Works !
A
!
A 76 - 180 Building Footprint
0 1 2 4 6 8 10
Kilometers
129
Figure 4.9: Sampling frequency for bromate at each monitoring location in 2007
!
! A
A
!
A
± !
A F
G !
A
!
A
!
A
! !
A !
A !
A ! !
A A
A
!
! ! ! !
A
!
A
A
A
A!! A A
A !
A !
A !
A
!
A !
A !
! A
A
!
A
! !
A !
A G
!
Source Site A !
A A !
! A
A !G
AF F G
F
!
F
G !
!
A
A !
A
A!A
!
A
! !
A !
A A
!
A !
A
! !
A
A F
G F
G
!
A ! G
F
Bromate A
Monitoring !
A
!
A !
A
Frequency F
G
!
A
Surface water
!
A
Yr_2008
F
G 1-5
!
A
F
G 6 - 10
!
A
F 11 - 25
G !
A
!
A
Groundwater
4.3. Monitoring Data Quality
Yr_2008
!
A 1-4
! 5 - 10 !
A
A
!
A
!
A 11 - 27
!
A
!
A
PWS !
A
Yr_2008 !
A
!
A 1 - 20
! Boundary of Source Site
A 21 - 75
Former Works !
A
!
A 76 - 180 Building Footprint
0 1 2 4 6 8 10
Kilometers
130
Figure 4.10: Sampling frequency for bromate at each monitoring location in 2008
4.4. Delineating the Bromate ‘Plume’ 131
cluded. Sampling frequencies were also reduced at most locations, although the aim was not to reduce
the frequencies until a full year of data at regular intervals had been obtained.
In June 2002, sampling frequencies were further amended with the aim to monitor at a frequency
of two-monthly or less with the exception of a small number of locations retained at monthly frequency
due to their strategic role in providing an early warning of change. Furthermore, in November 2002 the
sampling schedule was rationalised again due to funding constraints. Approximately 47 locations have
been retained for on-going and regular monitoring.
Figure 4.11: Annual average bromate concentrations at groundwater sampling locations in 2000.
4.4. Delineating the Bromate ‘Plume’
133
Figure 4.12: Annual average bromate concentrations at groundwater sampling locations in 2001.
4.4. Delineating the Bromate ‘Plume’
134
Figure 4.13: Annual average bromate concentrations at groundwater sampling locations in 2002.
4.4. Delineating the Bromate ‘Plume’
135
Figure 4.14: Annual average bromate concentrations at groundwater sampling locations in 2003.
4.4. Delineating the Bromate ‘Plume’
136
Figure 4.15: Annual average bromate concentrations at groundwater sampling locations in 2004.
4.4. Delineating the Bromate ‘Plume’
137
Figure 4.16: Annual average bromate concentrations at groundwater sampling locations in 2005.
4.4. Delineating the Bromate ‘Plume’
138
Figure 4.17: Annual average bromate concentrations at groundwater sampling locations in 2006.
4.4. Delineating the Bromate ‘Plume’
139
Figure 4.18: Annual average bromate concentrations at groundwater sampling locations in 2007.
4.4. Delineating the Bromate ‘Plume’
140
Figure 4.19: Annual average bromate concentrations at groundwater sampling locations in 2008.
4.4. Delineating the Bromate ‘Plume’ 141
SMD
80
80
40
40
0
0
78
76
74
6000
Hatfield swith off
Pump Test Start
Mean Monthly Br
Raw Br
Mean Monthly BrO3
Concentration (µg l-1)
2000
1
Bromate conc. / bromide conc.
0.8
0.6
0.4
0.2
Figure 4.20: Time series of bromate and bromide concentrations at selected locations between Sandridge
and Hatfield, soil moisture deficit, and monthly rainfall.
4.4. Delineating the Bromate ‘Plume’ 143
SMD
80
80
40
40
0
0
78
76
74
8000
Hatfield swith off
Pump Test Start
Mean Monthly Br
6000 Raw Br
Mean Monthly BrO3
Concentration (µg l-1)
4000
2000
0.8
0.6
0.4
0.2
Figure 4.21: Time series of bromate and bromide concentrations at selected locations between Sandridge
and Hatfield, soil moisture deficit, and monthly rainfall.
4.4. Delineating the Bromate ‘Plume’ 144
SMD
80
80
40
40
0
0
Raw BrO3
6000
4000
2000
0.8
0.6
0.4
0.2
Figure 4.22: Time series of bromate and bromide concentrations at selected locations between Sandridge
and Hatfield, soil moisture deficit, and monthly rainfall.
4.4. Delineating the Bromate ‘Plume’ 145
deviation is not supported by concentrations above 50 µg l−1 which are consistently monitored at loca-
tion 065, location 372 and location 378.
The locations 067, 065 and 166 provide good time series data over the period 2000 to 2008 (Fig-
ure 4.23 to 4.23). Locations 067 and 065 both show a rise in bromate (and bromide) concentrations in
2002 compared to 2000 and 2001 concentrations. Bromate concentrations then remain fairly stable until
the end of 2005, after which concentrations at location 067 fall to levels comparable to 2000 and 2001
(with a particularly low concentration in October 2007). Locations 166 shows large fluctuations in its
time series, However maximum concentrations remain at similar levels between 2000 and 2008, and no
clear trend is discernible.
At location 068, bromate concentrations have decreased from between 10 and 25 µg l−1 over 2001,
to around 5 µg l−1 from 2002 to 2008. Although there is a gap in sample results between October 2002
and October 2005, the following observations can be made regarding trends in bromate concentrations
at locations around the Quarry area:
4.4. Delineating the Bromate ‘Plume’ 146
SMD
80
80
40
40
0
0
76
72
Orcharge Garage (Ref. 028) 68
Hatfield Quarry WM9 (Ref. 065)
Hatfield swith off
Pump Test Start
600 Mean Monthly Br
Raw Br
Mean Monthly BrO3
Raw BrO3
Concentration (µg l-1)
400
200
0.8
0.6
0.4
0.2
Figure 4.23: Time series of bromate and bromide concentrations at selected locations in the Hatfield
Quarry area, soil moisture deficit, and monthly rainfall.
4.4. Delineating the Bromate ‘Plume’ 147
SMD
80
80
40
40
0
0
76
72
Orcharge Garage (Ref. 028) 68
Hatfield Quarry WM12 (Ref. 067)
Hatfield swith off
Pump Test Start
5000 Mean Monthly Br
Raw Br
Mean Monthly BrO3
4000
Raw BrO3
Concentration (µg l-1)
3000
2000
1000
0.8
0.6
0.4
0.2
Figure 4.24: Time series of bromate and bromide concentrations at selected locations in the Hatfield
Quarry area, soil moisture deficit, and monthly rainfall.
4.4. Delineating the Bromate ‘Plume’ 148
SMD
80
80
40
40
0
0
76
72
2000
1000
0.8
0.6
0.4
0.2
Figure 4.25: Time series of bromate and bromide concentrations at selected locations in the Hatfield
Quarry area, soil moisture deficit, and monthly rainfall.
4.4. Delineating the Bromate ‘Plume’ 149
• Locations 061 and 064 show an apparent decreasing trend in bromate concentrations
• Locations 062 and 066 show generally stable trend in bromate concentrations
Since 2005:
Additionally, locations 059, 060, 162 and 199 to the south of Hatfield Quarry define the southern
margin of the bromate contamination, and locations 007, 380, 381 and 373 to the north of Hatfield
Quarry define the northern margin of the bromate contamination. Locations 060 and 199 show only
sporadic occurrences of bromate above the MDL. Location 059 shows concentrations below MDL or
just above (<1 µg l−1 ): the incidences of concentrations above the MDL become more frequent after
June 2003, and three successive results between 2 and 5 µg l−1 occur in the latter half of 2006. Bromide
concentrations show no obvious trend between 2001 and 2008. In contrast, location 162 shows variable
concentrations up to ∼ 12 µg l−1 , with slight rising trend observable until the end of 2006. Subsequent
concentrations have been below the MDL. Bromide concentrations have also decreased over 2007 and
2008.
Around the northern margin, at location 380, bromate concentrations have decreased from between
1 and 3 µg l−1 between late 2005 and early 2007, to below the MDL over 2007 and 2008. Bromate
concentrations are much more variable at location 381, fluctuating between 1 and 5 µg l−1 (although
one concentration of 9.3 µg l−1 in July 2008) with no discernible trend in bromate (or bromide) concen-
trations.
December 2001. The observations have been the subject of much discussion, and the Environment
Agency concluded that leakage from shallow depth (via the borehole annulus) was impacting on borehole
water quality (Atkins, 2006). Location 402 (Comet Way BH5) shows a rising trend of bromate (and
bromide) concentrations from January 2006 and December 2008, with concentrations increasing from
∼ 400 µg l−1 to ∼700 µg l−1 . Location 160 shows a declining trend in bromate concentrations, from
∼ 800 µg l−1 in September 2001 to ∼ 500 µg l−1 in July and August 2002, although concentrations
then increase to ∼700 µg l−1 in September and October 2002. No samples were taken again until May
2005, when concentrations were recorded below MDL until September 2005.
To the north, at location 378 bromate concentrations show a declining trend from ∼ 140 µg l−1 at
the end of 2005 to ∼90 µg l−1 at the end of 2008. At location 379 bromate concentrations are generally
around 40 to 60 µg l−1 , although a slight declining trend appears to occur over 2006, 2007 and 2008.
Bromide concentrations mirror bromate concentration trends for both of these boreholes. The northern
margin is defined by locations 016/162 (Old Cottage, Green Lane, Old BH/New BH), 375 and 376.
Location 016/162 shows only a few isolated results above the MDL. Location 375 yields concentrations
around 1 µg l−1 , with a peak of 3 to 4 µg l−1 in Autumn 2006.
The southern margin is defined by location 235 (Carter’s Pond BH), location 142 and location 195.
Locations 142 and 195 are almost always below the MDL. Time series for location 235 (Carter’s Pond)
shows bromate below the MDL, with the exception of intermittent concentrations of 1 to 2 µg l−1 in late
2001 and early 2002. Further south, locations 049 (Brand’s Nursery BH), location 006, 014 and 015 are
typically below MDL, but show intermittent concentrations above the MDL.
Slight rising trends in bromate concentrations are observed at locations 191 (Mill Green
BH) and 265 (Park Street BH) to the east towards the Lea Valley. Bromate concentrations rise from
∼2 to 4 µg l−1 in 2001 and 2002 to ∼8 to 10 µg l−1 in 2004 to 2008 (with lower concentrations of
∼ 2 to 4 µg l−1 in late 2006 and early 2008) for 191. However, there are a number of results below MDL
at 191. The time series for 265 is very variable, with concentrations fluctuating considerably. However,
there appear to be two parallel trend lines: one from ∼2 to 4 µg l−1 in 2002 to ∼18 to 22 µg l−1 in 2008,
and another from ∼40 to 45 µg l−1 in 2002 to ∼50 to 60 µg l−1 in 2007, and a fall in concentrations to
∼45 µg l−1 in 2008.
SMD
80
80
40
40
0
0
85
75
Orcharge Garage (Ref. 028)
Business Park (Ref. 002) 70
65
60
Hatfield swith off
2000 Pump Test Start
Mean Monthly Br
Raw Br
1600 Mean Monthly BrO3
Concentration (µg l-1)
Raw BrO3
1200
800
400
0.8
0.6
0.4
0.2
Figure 4.26: Time series of bromate and bromide concentrations at selected locations in the Hatfield
area, soil moisture deficit, and monthly rainfall.
4.4. Delineating the Bromate ‘Plume’ 152
SMD
80
80
40
40
0
0
84
Raw BrO3
1500
1000
500
0.8
0.6
0.4
0.2
Figure 4.27: Time series of bromate and bromide concentrations at selected locations in the Hatfield
area, soil moisture deficit, and monthly rainfall.
4.4. Delineating the Bromate ‘Plume’ 153
200
Monthly Rainfall (mm) Comet Way BH5 (Ref. 402)
Rothamsted Station
SMD
80
80
40
40
0
0
85
Jan-00 Jan-02 Jan-04 Jan-06 Jan-08 Jan-10
Raw BrO3
800
400
0.8
0.6
0.4
0.2
Figure 4.28: Time series of bromate and bromide concentrations at selected locations in the Hatfield
area, soil moisture deficit, and monthly rainfall.
4.4. Delineating the Bromate ‘Plume’ 154
160
SMD
80
80
40
40
0
0
85
Jan-00 Jan-02 Jan-04 Jan-06 Jan-08 Jan-10
80
60
200 55
Mean Monthly Br
120 Raw Br
Mean Monthly BrO3
Raw BrO3
80
40
0.8
0.6
0.4
0.2
Figure 4.29: Time series of bromate and bromide concentrations at selected locations in the Hatfield
area, soil moisture deficit, and monthly rainfall.
4.4. Delineating the Bromate ‘Plume’ 155
200
Monthly Rainfall (mm) Park Street (Ref. 265)
Rothamsted Station
SMD
80
80
40
40
0
0
90
70
Orcharge Garage (Ref. 028)
Park Street (Ref. 265) 60
50
120
80
40
0.8
0.6
0.4
0.2
Figure 4.30: Time series of bromate and bromide concentrations at selected locations in the Hatfield
area, soil moisture deficit, and monthly rainfall.
4.4. Delineating the Bromate ‘Plume’ 156
Between the Hatfield area and Essendon, at location 266 (Hill End Farm BH), there have been
two occurrences of bromate concentrations around 75 µg l−1 , but all other concentrations have been
below the MDL. These two high results lead to mean annual concentrations between 30 and 40 µg l−1
in 2003 and 2004. These results may be anomalous. In support of this, locations 262, 263 and 364,
in close proximity to 266, all show concentrations below MDL. This is despite these locations be-
ing directly between bromate in the Hatfield area and Location 143 (Essendon P.S.) where bromate
concentrations have consistently been above 10 µg l−1 . Interestingly, bromate is also not observed
at 144 (Water Hall P.S.), on the southern side of the River Lea, but is recorded at certain locations
to the north-east on the northern side of the River Lea at Southfield Wood Landfill boreholes (loca-
tions 329 to 331, 364 to 367 and 369 to 370). Bromate concentrations around 10 to 20 µg l−1 are
recorded at selected locations, generally in the central to north-eastern parts of this area. Additionally,
location 404 shows annual average bromate concentrations around 2 to 5 µg l−1 , although approximately
half of the samples are below MDL.
Bromate concentrations appear to have increased at location 089 (Holly Cottage BH), situated to the
north-east of Southfield Wood Landfill, but on the southern side of the River Lea. Bromate concentrations
generally rise from below MDL in 2000, to between 5 and 7 µg l−1 between October 2007 and July 2008.
This increase is accompanied by an increase in bromide concentrations. A bromate sample in October
2008 shows bromate concentrations return to 2 µg l−1 ; bromide concentrations remained on a rising
trend. Continuing on east along the Lea, locations 305 to 307 are generally below MDL, and always less
than 1 µg l−1 .
Location 005 (Hatfield London Country Club BH) appears to mark the southern margin; bromate
has been recorded on an intermittent basis (in approx two-thirds of the samples taken) throughout the
monitoring period, up to a maximum of 13.0 µg l−1 . There are a number of locations between 005 and
the River Lee that show concentrations below the MDL (Locations 405 to 409). These were monitored
once in 2007. Location 005 is at the end of an inferred conduit route, and locations 405 to 409 are to the
east of this.
The locations along the Northern New River define the easterly extent of the bromate contamination.
Chadwell Spring shows bromate concentrations intermittently above MDL. Chadwell Spring feeds into
the New River. In general, the most northerly of the Northern New River wells, locations 295 (Amwell
End P.S.) and 298 (Broadmeads P.S.), show bromate concentrations below the MDL, although sporadic
occurrences of bromate above the MDL have occurred, particularly since 2004. Annual average concen-
trations tend to be highest in the central and southern parts of the wellfield: Hoddesdon, Broxbourne,
Turnford and Amwell Marsh.
There are relatively few additional sampling locations to the west of the the River Lee - New River
Loop. Location 354 (Van Hage Nurseries BH) shows bromate concentrations below MDL. In 2003,
locations 304 and 309, and locations 293 and 394 and locations 312 to 316 were sampled once, and
showed bromate concentrations below MDL.
4.4. Delineating the Bromate ‘Plume’ 157
SMD
80
80
40
40
0
0
120
80
40
0.8
0.6
0.4
0.2
Figure 4.31: Time series of bromate and bromide concentrations at selected locations in the Lea Valley,
soil moisture deficit, and monthly rainfall.
4.5. Bromate-bromide ratios 158
401 - 600
601 - 2000
2001 - 1000
1001 - 2000
2001 - 4000
4001 - 20000
20001 - 100000
100001 - 500000
0 1 2 3 4 5
Kilometers
© Crown Copyright/database right 2008. An Ordnance Survey/EDINA supplied service.
159
Groundwater
Bromide 00-08 F
G
0 20 40 60 80 100 Meters
25 - 50
51 - 100
101 - 150
151 - 200 F
F
G GG
F
201 + F
G
Surface water F
G
0 0.5 1GF
Bromide 00-08 F
F
G
Kilometers G
F 25 - 50
G
F 51 - 100
G
F 101 - 150
G
G 151 - 200
F
F 201+
G F
G
Sample used
to calculate
Bromide background
concentration F
G
F
G
F
G G F
F G
F
G
G
F F
GG
F F
G
F
G G
F G
F
4.5. Bromate-bromide ratios
F
G
F
G F
G
F
G
F
G
F
G F
G
F
G F
G
F
G
F
G
F
G
0 1 2 3 4 5
Kilometers
© Crown Copyright/database right 2008. An Ordnance Survey/EDINA supplied service.
Geological Map Data © NERC 2008.
160
Figure 4.33: Bromide concentrations at locations where bromate concentrations are less than MDL.
4.5. Bromate-bromide ratios 161
2.0
Surface water
Groundwater - PWS
1.6 Groundwater - Chalk
Groundwater - Gravel
Bromate/bromide ratio
1.2
0.8
0.4
0.0
0.8
0.4
0.8
0.6
0.4
0.2
0.4
0.2
Figure 4.34: Bromate/Bromide ratio variation with bromate concentration for groundwater and surface
water samples
4.6. Bromate concentrations and water levels 162
Table 4.3: Regression statistics for the response of bromide concentration to bromate concentration and
chloride concentration.
The spatial distribution of bromate-bromide ratio is shown in Figure 4.35. In general, the higher
ratios follow the distribution of bromate contamination. If locations with bromate concentrations less
than 1 µg l−1 ) are excluded from the plot, there is a noticeable cluster of locations with high ratios in the
northern part of the St Leonard’s Court source site, and following the path of the core bromate ‘plume’
to Hatfield Quarry and the Hatfield area.
0.81 - 1.60
0.81 - 1.60
0.00 - 0.10
0.11 - 0.30
0.31 - 0.50
0.51 - 0.80
0.81 - 1.60
Figure 4.35: Spatial distribution of mean annual bromate/bromide ratio 2000 to 2008.
4.7. Summary and Conclusions
164
Figure 4.36: Percentage of samples of bromate concentrations for which there are accompanying water level measurements for each location
!
(
(
!
^
_ (
!
_^_
^ (
!
_
^ (
!
!
( !
(
^ _
^
_^_ _
^ !
(
(
! _
^
(
!
(
!
_
^
(
!
±
_
^ (
! (
!
_
^
(
!
0 20 40 60 80 100 Meters
(
!
_
^
_
^
(
!
Regression:
Bromate conc. versus
water level (m OD) 0 0.5 1 _
^
Negative Regression Kilometers
P-value < 0.05
_
^ 0% - 25%
_
^ 26% - 50%
_
^ 51% - 75%
_
^ 76% - 100%
Negative Regression
P-value > 0.05
(
!!
( (
!
(
! 0% - 25% _
^^_ _^ ^
_
^ (
! (
! _!
_!(^ (
!
( 26% - 50% (
! _
^ _
^
_
^ !
( !
(
(
_
^
!!
(
( 51% - 75%
! _
^ _
^ (
!
(
! (
!
_
^ _
^
(
! 76% - 100% _
^ (
!
Positive Regression
P-value < 0.05
4.7. Summary and Conclusions
_
^ 0% - 25%
_
^ 26% - 50%
_
^ 51% - 75%
_
^ 76% - 100%
Positive Regression
P-value > 0.05
(
! 0% - 25%
!
( 26% - 50%
( 51% - 75%
!
(
! 76% - 100%
0 1 2 3 4 5
Kilometers
© Crown Copyright/database right 2008. An Ordnance Survey/EDINA supplied service.
Geological Map Data © NERC 2008.
165
Figure 4.37: Regression relationship for the response of bromate concentration to water level. Percentages refer to the amount of variation explained by the regression (R2
value)
4.7. Summary and Conclusions 166
revised conceptual understanding of flow and transport in the Hertfordshire Chalk has allowed a new
interpretation of the spatial distribution and evolution of bromate and bromide within the catchment to
be developed between 2000 and 2008. However, the interpretation of the spatial and temporal evolution
of bromate and bromide within the catchment is hampered by a number of inadequacies in the available
monitoring data:
• Monitoring data are available for a relatively short period of time (a maximum of 8 years continu-
ous data) in relation to the likely timescale of bromate contamination within the catchment;
• Monitoring frequency varies considerably between locations, and varies over time at individual
locations, which makes trends difficult to identify with confidence;
• The strong seasonal influences within the time series make trends difficult to discern;
• The data available are generally not depth-specific so that vertical distribution of bromate contam-
ination cannot be investigated;
• The sampling results refer to (mobile) fissure water and there are no data available for (immobile)
matrix porewater which is required to characterise the double-porosity behaviour and determine
the long-term evolution of bromate contamination within the catchment.
167
Chapter 5
• To describe and quantify the distribution of bromate at the source site through collation and de-
scription of site investigation and monitoring data;
• To develop alternative conceptual scenarios for bromate release to groundwater and quantify these
as ‘source terms’;
• To use the available monitoring data to constrain the potential source terms.
Figure 5.1: Location of the source site in Sandridge, Hertfordshire. Formerly the Steetly chemical works,
now the St Leonard’s Court residential development.
5.4. Chronology and scope of investigations 169
• Original records from St Albans District Council (the local authority); and
• Additional information collected by the Environment Agency in response to a public request for
information conducted in 2001.
An interview with a former worker at the chemical works, conducted by the Environment Agency in
August 2001, provided details on the operational activities (EA, 2005). Aerial photograph taken in 1971.
Figure 5.2: Location of former process areas of the Steetly Chemical Works. Based on Atkins (2002)
interpretation of historical plans, aerial photographs, and the interview with a former employee of the
works. Aerial photograph taken in 1971.
Table 5.1: Chronology and scope of site investigations and monitoring at the source site
Date of Date of
Company Scope of investigation Associated Reports Client
site work report
Borehole C1 drilled to 5.4m deep. Core samples tested for Field report for drilling of borehole C1 at House
Jan 1985 Chemfix Mar 1985 Crest
moisture content and bromide. Lane, Sandridge on 21/22 January 1985
Southern 51 shallow boreholes, depth generally approx 1.5m. Hand-augered borehole logs at Sandridge, Herts
Mar 1985 Mar 1985 Chemfix
Testing Analysis for bromide. for Chemfix International Ltd
St Albans
5 boreholes to depths of approx 12m. Soil and Site Investigation at St Leonards Court, City and
Aug 2000 Komex Oct 2000
groundwater samples. Analysis for bromide and bromate. Sandridge, St Albans. District
Council
Figure 5.3: Borehole locations from investigations 1983-1985 (STATS, 1983a,b,c, 1984; Chemfix,
1985c) and 2000-2001 (Komex, 2000; Atkins, 2002). For locations from 1983-1985, numbers in square
brackets indicate date of drilling.
5.4. Chronology and scope of investigations 173
Figure 5.4: Trial hole locations from investigations in 1985 (Chemfix, 1985c)
5.5. Site Geology and Hydrogeology 174
Table 5.2: Geological strata encountered at the source site. Based on Komex (2000) and Atkins (2002)
‘Putty’ Chalk
variable Structureless, weathered Chalk with occasional flints.
(Upper Chalk)
‘Blocky’ Chalk
>10 m Chalk with horizontal and vertical fractures.
(Upper Chalk)
Groundwater levels varied between +78.78 m OD and 81.15 m OD during the 2001 investigation
(Atkins, 2002). Borehole logs show that groundwater appears to be semi-confined by the low perme-
ability ’putty chalk’: there is generally a rise in groundwater rest levels compared to strike levels. Piezo-
metric contours (Figure 5.5) indicate a south-easterly flow direction, with a hydraulic gradient across
site of 0.0042 (Atkins, 2002). The local hydraulic gradient is lower: 0.0028 in south-easterly direction
(Atkins, 2002). Vertical groundwater elevation contours on section along flow direction indicate flow is
principally horizontal (Figure 5.6).
Figure 5.5: Piezometry at the St Leonard’s Court site November 2001. From Atkins (2002)
5.6. Contaminant Distribution
176
Figure 5.6: Cross-section parallel to groundwater flow direction. From Atkins (2002)
5.6. Contaminant Distribution 177
Fluvio-glacial
deposits
Chalk
Figure 5.7: Spatial distribution of the bromide contamination based on investigations undertaken be-
tween 1983-1985.
The highest bromide concentrations (>1000 mg kg−1 ) occur in soil samples from boreholes in
the vicinity of the former ‘solid bromate handling’ and ‘bulk bromine storage’ areas, and close to the
sump in the ‘non-bromate production’ area (Figure 5.7). Locations in the southern and eastern areas of
the site, corresponding to the non-process areas of the site showed much lower bromide concentrations
(<200 mg kg−1 ).
5.6. Contaminant Distribution 178
Concentration-depth profiles for these boreholes (Figure 5.8) indicate highest bromide concentra-
tions in the Made Ground and in the Putty Chalk, with generally lower concentrations in the fluvio-glacial
deposits (clayey gravels).
The results of this sampling was used as a basis for the excavation of between 0.75 m and 1.5 m
of the top layer of soil over much of the site as part of remediation carried out between 1985 and 1986
(Roberts, 2001). It is unclear whether any verification samples were submitted, or whether there were
significant alterations to these proposals.
Clay
83 83 83 Sand
Chalk
82 82 82 Monitoring well
screened interval
81 81 81
Calculated Porewater
80 C1-[85]
80 80 Porewater Bromide Conc.
Jan-85
BH1-[84]
79 79 79 Porewater Bromide Conc.
Mar-84
BH3-[84]
Porewater Bromide Conc.
78 78 78 Mar-84
BH2-[84]
82 82 82
8000 20
5.6. Contaminant Distribution
81 81 81
80 80 80 6000 15
79 79 79
4000 10
78 78 78
Water Level (mAOD)
77 77 77 2000 5
76 76 76
75 0 0
75 75
0 4000 8000 12000 16000 Jan-84 Apr-84 Jul-84 Oct-84 Jan-85
bromide concentration (mg/l)
179
Figure 5.8: Depth profiles of porewater bromide compared to pumped groundwater concentrations for boreholes from investigations 1983-1985.
5.6. Contaminant Distribution 180
Fluvio-glacial
deposits
Chalk
Figure 5.9: spatial distribution of bromide (as mg kg−1 ) based on investigations undertaken between
2000 and 2001.
5.6. Contaminant Distribution 181
Fluvio-glacial
deposits
Chalk
Figure 5.10: spatial distribution of bromate (as mg kg−1 ) based on investigations undertaken between
2000 and 2001.
Borehole 214 Porewater concentrations
214
84 Bromate Bromide
84 84
83
83 83
82 82 82
81 81 81
80 80 80
Made Ground
Clay
79 79 79
Sand
78 78 78
Monitoring well
screened interval
77 77 77
Pumped Groundwater Conc.
Nov-01 0 5 10 15 20 25 0 5 10 15 20 25
Porewater Bromate Conc.
Nov-01 bromate concentration (mg/l) bromide concentration (mg/l)
Soil Bromate Conc.
Nov-01
0.0020 16
78 78
0.0016 12
76 76
0.0012 8
Water Level (mAOD)
74 74
Bromide concentration (mg/l)
72 72
0.0004 0
May-01 May-02 May-03 May-04 May-05 May-01 May-02 May-03 May-04 May-05
182
Figure 5.11: Depth profiles of porewater bromate and bromide compared to pumped groundwater concentrations for Borehole 214 from 2001 investigation (Atkins, 2002).
Borehole 215
Porewater concentrations
215
84
Bromate Bromide
85 85
83
84 84
82 83 83
82 82
81
81 81
80
Made Ground
80 80
Clay
Sand
79
Monitoring well 78 78 78
screened interval
77 77 77
Pumped Groundwater Conc.
Nov-01 0 5 10 15 20 25 0 20 40 60 80
Porewater Bromate Conc.
Nov-01 bromate concentration (mg/l) bromide concentration (mg/l)
Soil Bromate Conc.
Nov-01
Water Level
80 200 80
1
78 78
1 150
76 76
0 100
Water Level (mAOD)
74 74
Bromide concentration (mg/l)
72 72
0 0
May-01 May-02 May-03 May-04 May-05 May-01 May-02 May-03 May-04 May-05
183
Figure 5.12: Depth profiles of porewater bromate and bromide compared to pumped groundwater concentrations for Borehole 215 from 2001 investigation (Atkins, 2002).
Borehole 216
Porewater concentrations
216
84 Bromate Bromide
84 84
83
83 83
82
82 82
81 81 81
80 80 80
79 79 79
78 78 78
77 77 77
Made Ground 76 76 76
Clay
75 75 75
Sand
Water Level
80 80
4
3
78 78
3
2
76 76
2
Water Level (mAOD)
74 1 74
Bromide concentration (mg/l)
72 72
0 0
May-01 May-02 May-03 May-04 May-05 May-06 May-01 May-02 May-03 May-04 May-05 May-06
184
Figure 5.13: Depth profiles of porewater bromate and bromide compared to pumped groundwater concentrations for Borehole 216 from 2001 investigation (Atkins, 2002).
Borehole 217
217 Porewater concentrations
85
Bromate Bromide
84 84 84
83 83 83
82 82 82
81 81 81
80 80 80
79 79 79
78 78 78
77 77 77
Made Ground 76 76
76
Clay
75 75 75
Sand
73 73 73
Monitoring well
screened interval
72 72 72
71 71 71
Pumped Groundwater Conc.
Nov-01 0 2 4 6 8 10 0 20 40 60 80
Porewater Bromate Conc.
Nov-01 bromate concentration (mg/l) bromide concentration (mg/l)
Soil Bromate Conc.
Nov-01
Water Level
80 1.6 80
1.2
78 78
1.2
0.8
76 76
0.8
Water Level (mAOD)
74 74
0.4
Bromide concentration (mg/l)
72 72
0.0 0.0
May-01 May-02 May-03 May-04 May-05 May-06 May-07 May-01 May-02 May-03 May-04 May-05 May-06 May-07
185
Figure 5.14: Depth profiles of porewater bromate and bromide compared to pumped groundwater concentrations for Borehole 217 from 2001 investigation (Atkins, 2002).
Borehole 218 218
85 Porewater concentrations
Bromate Bromide
84 84 84
83 83 83
82 82 82
81 81 81
80 80 80
Made Ground
Clay
Sand 79 79 79
Monitoring well 78 78 78
screened interval
77 77 77
Pumped Groundwater Conc.
Nov-01 0 2 4 6 8 0 4 8 12 16
Porewater Bromate Conc.
Nov-01 bromate concentration (mg/l) bromide concentration (mg/l)
Soil Bromate Conc.
Nov-01
Water Level
8
6 81 81
4 80 80
6
Water Level (mAOD)
2
Bromide concentration (mg/l)
0 4
May-01 May-02 May-03 May-04 May-05 May-01 May-02 May-03 May-04 May-05
186
Figure 5.15: Depth profiles of porewater bromate and bromide compared to pumped groundwater concentrations for Borehole 218 from 2001 investigation (Atkins, 2002).
Borehole 219 219 BH Log
85
Porewater concentrations
Bromate Bromide
84 84
84
83 83 83
82 82 82
81 81 81
80 80 80
Made Ground
Clay
Sand 79 79 79
Monitoring well
78 78 78
screened interval
77 77 77
Pumped Groundwater Conc.
Nov-01 0 100 200 300 400 500 0 100 200 300 400 500
Porewater Bromate Conc.
Nov-01 bromate concentration (mg/l) bromide concentration (mg/l)
Soil Bromate Conc.
Nov-01
Water Level 20
16
81 18 81
12
16
8
80 14 80
Water Level (mAOD)
0 10
May-01 May-02 May-03 May-04 May-05 May-01 May-02 May-03 May-04 May-05
187
Figure 5.16: Depth profiles of porewater bromate and bromide compared to pumped groundwater concentrations for Borehole 219 from 2001 investigation (Atkins, 2002).
Borehole 220
Porewater concentrations
220
85
Bromate Bromide
85 85
84 84 84
83 83 83
82 82 82
81 81 81
Made Ground 80 80 80
Clay
Sand
Monitoring well
78 78 78
screened interval
77 77 77
Pumped Groundwater Conc.
Nov-01 0 2 4 6 8 10 0 20 40 60 80
Porewater Bromate Conc.
Nov-01 bromate concentration (mg/l) bromide concentration (mg/l)
Soil Bromate Conc.
Nov-01
Water Level 7 80 80
20
6 78
78
5 18
76 76
4
Water Level (mAOD)
74 16 74
Bromide concentration (mg/l)
2 14
May-01 May-02 May-03 May-04 May-05 May-01 May-02 May-03 May-04 May-05
188
Figure 5.17: Depth profiles of porewater bromate and bromide compared to pumped groundwater concentrations for Borehole 220 from 2001 investigation (Atkins, 2002).
Borehole 221 221 BH Log
85
Porewater concentrations
Bromate Bromide
84 84
84
83
83 83
82 82 82
81 81 81
80 80 80
Made Ground
Clay
Sand 79 79 79
78 78 78
Monitoring well
screened interval
77 77 77
Pumped Groundwater Conc.
Nov-01 0 400 800 1200 1600 2000 0 200 400 600 800 1000
Porewater Bromate Conc.
Nov-01 bromate concentration (mg/l) bromide concentration (mg/l)
Soil Bromate Conc.
Nov-01
Water Level
400
160
81 81
300
120
200
80 80
Water Level (mAOD)
80
Bromide concentration (mg/l)
0 40
May-01 May-02 May-03 May-04 May-05 May-01 May-02 May-03 May-04 May-05
189
Figure 5.18: Depth profiles of porewater bromate and bromide compared to pumped groundwater concentrations for Borehole 221 from 2001 investigation (Atkins, 2002).
Borehole 222 222 Porewater concentrations
85
Bromate Bromide
84 84
84
83 83
83
82 82 82
81 81 81
80 80 80
Made Ground
Clay
Sand 79 79 79
78 78 78
Monitoring well
screened interval
77 77 77
Pumped Groundwater Conc.
Nov-01 0 50 100 150 200 250 0 100 200 300 400
Porewater Bromate Conc.
Nov-01 76 bromate concentration (mg/l) bromide concentration (mg/l)
Soil Bromate Conc.
Nov-01
Water Level 81 81
80 60
81 81
60 80 50 80
80 80
Water Level (mAOD)
40 40
Bromide concentration (mg/l)
Figure 5.19: Depth profiles of porewater bromate and bromide compared to pumped groundwater concentrations for Borehole 222 from 2001 investigation (Atkins, 2002).
Borehole 223 223 Porewater concentrations
85
84 Bromate Bromide
84 84
83 83 83
82 82 82
81 81 81
80 80 80
79 79 79
78 78 78
77 77 77
Made Ground
Clay 76 76 76
Sand
72 72 72
Pumped Groundwater Conc.
Nov-01 0 100 200 300 400 500 0 100 200 300 400 500
Porewater Bromate Conc.
Nov-01 bromate concentration (mg/l) bromide concentration (mg/l)
Soil Bromate Conc.
Nov-01
Water Level
81 81
10
8
81 81
80 80
4
Water Level (mAOD)
6
Bromide concentration (mg/l)
0 4
May-01 May-02 May-03 May-04 May-05 May-01 May-02 May-03 May-04 May-05
191
Figure 5.20: Depth profiles of porewater bromate and bromide compared to pumped groundwater concentrations for Borehole 223 from 2001 investigation (Atkins, 2002).
Borehole 225
Porewater concentrations
225 BH Log
84 84 Bromate 84 Bromide
82 82
80 80 80
78 78
76 76 76
Made Ground 74 74
Clay
Monitoring well 70 70
screened interval
68 68 68
Pumped Groundwater Conc.
Nov-01 0 0.05 0.1 0.15 0.2 0.25 0 100 200 300 400 500
Porewater Bromate Conc.
Nov-01 bromate concentration (mg/l) bromide concentration (mg/l)
Soil Bromate Conc.
Nov-01
Water Level
80 80
0.006
200
78 78
0.004
76 76
100
Water Level (mAOD)
0.002
Bromide concentration (mg/l)
0.000 0
May-01 May-03 May-05 May-07 May-01 May-03 May-05 May-07
192
Figure 5.21: Depth profiles of porewater bromate and bromide compared to pumped groundwater concentrations for Borehole 225 from 2001 investigation (Atkins, 2002).
5.6. Contaminant Distribution 193
therefore possible that bromate and/or bromide contamination could extend further to the northern part
of the site. Low bromate concentrations (<0.5 mg kg−1 ) were encountered in the southern and eastern
part of the site. Bromate concentrations were below detection limits of 0.10 mg kg−1 in all but one
of the samples tested during the August 2000 investigation (Komex, 2000). The samples included the
fluvio-glacial deposits and Chalk strata, at a range of depths up to 11.7 m bgl. It is possible that bromate
was not detected because the locations were sited away from the bromate production and handling areas.
Clear concentration-depth patterns are difficult to discern for the bromate depth profiles. However, in
a number of boreholes where elevated bromate concentrations are encountered, higher concentrations
occur in the Chalk (putty chalk) compared to concentrations in the shallower fluvio-glacial deposits
(clays and gravels).
Bromate conc. (mg kg−1 ) = 0.36 × Bromide conc. (mg kg−1 ) + 32.1 (mg kg−1 )
120
Soil Bromate (mg kg-1)
80 Fit Results
Fit 1: Linear
Equation Y = 0.3574023353 * X + 32.0602463
Number of data points used = 7
Average X = 90.2857
40
Average Y = 64.3286
Residual sum of squares = 1077.19
Regression sum of squares = 7955.88
Coef of determination, R-squared = 0.88075
Residual mean square, sigma-hat-sq'd = 215.439
0
Figure 5.22: Relationship between soil bromate and soil bromide concentrations based on soil samples
from the 2001 site investigation (Atkins, 2002).
5.6. Contaminant Distribution 194
4.0E+005
220 210500 210500 210500
3.6E+005 210500
3.2E+005 079 221
219
2.8E+005 214 216
222
2.4E+005 217
218
2.0E+005 080
1.6E+005 210450 215 210450 210450 210450
223
1.2E+005 081 Flux Flux Flux
Flux 082
8.0E+004 plane plane plane
4.0E+004
plane
083
0.0E+000
210400 210400 210400 210400
517050 517100 517050 517100 517050 517100 517050 517100
Bromide concentration Groundwater Bromide Groundwater Bromide Groundwater Bromide Groundwater Bromide
(microgram per litre) Oct 2001 - Dec 2001 Jan 2002 - Mar 2002 Apr 2002 - Jun 2002 Jul 2002 - Sep 2002
4.0E+005
220 210500 210500 210500
3.6E+005 210500
5.6. Contaminant Distribution
Figure 5.23: Groundwater bromate and bromide contours at the ‘source zone’ based on samples taken in 2001 and 2002.
5.6. Contaminant Distribution 196
were approximately five times greater than sampled groundwater concentrations at the same location
in BH2-[84] and BH3-[84], but they were similar in BH1-[84].
Bromide concentrations at borehole 225 compared to borehole B1-[85] (Figure 5.24) which was
located in approximately the same location, indicate slightly higher concentrations in 2002 compared to
1985/1986. However, concentrations decline substantially into 2004 and remain low into 2007.
120000 20
6
200000
100000 15
80000 10
225 Bromide
225 Water Level
0 0 40000 0
May-01 May-02 May-03 May-04 May-05 May-06 May-07 May-08 May-84 May-85 May-86 May-87
84 84 Porewater Bromide
225
84 Bromide
Groundwater Bromide conc.
82 Porewater Bromide conc.
80 80
80
5.6. Contaminant Distribution
78
76 76
76
74
Sample elevation (mAOD)
72 72
70
68 68
68
0 100000 200000 300000 400000 500000 0 200000 400000 600000 800000
bromide concentration (µg/l) Bromide concentration (µg/l)
197
Figure 5.24: Depth profiles of porewater bromide compared to pumped groundwater concentrations for Borehole B1 from the 1985 investigation (Chemfix, 1985a) and
Borehole 225 from 2001 investigation (Atkins, 2002) which are believed to have been located in similar positions.
5.6. Contaminant Distribution 198
Figure 5.25: Relationship between bromide and bromate concentration in groundwater samples from the
monitoring data between 2000 and 2008 for locations 079 to 083 and locations 214 to 223.
bromate and bromide is seen over a wider area in 2001, followed by a rapid decline. Bromate relation-
ship to water levels have been assessed in Section 4.6. At the source site, the regression relationships for
the response of bromate concentration to water level are only statistically significant (P<0.05) for three
locations out of the six locations with sufficient data to undergo statistical analysis (Figure 4.37).
225
B1[85]
028
017
025
018
226
020
220
079 221
219
BH1[84]214216 222
217
218
C1[85]080 227
5.6. Contaminant Distribution
215
BH3[84] 223
081 019
082
083
0 50
Meters 225
B1[85]
Figure 5.26: Groundwater monitoring locations in the vicinity of the source site that have been sampled for bromide concentrations between 1983 and 1987 and between 2000
and 2008.
300000 8000
225 028
019
100000 4000
028
017
025
0
KEY
Jan-83 Jan-86 Jan-89 Jan-92 Jan-95 Jan-98 Jan-01 Jan-04 Jan-07
Average Bromide 2000-2002 018
not present in 2000-02
0 - 10 mg/l 226
10 - 100 mg/l
100 - 500 mg/l 020
Boundary of
former factory site
227
6000 200 017
019
018
020
160
025
120
8000
228
5.6. Contaminant Distribution
227 4000
226 80
224
6000
40
Bromide concentration (µg/l)
0
4000
2000
Bromide concentration (µg/l) May-00 May-01 May-02 May-03 May-04 May-05 May-06 May-07
228
0
0
May-01 May-02 May-03 May-04 May-05 May-06 May-07 Jan-83 Jan-86 Jan-89 Jan-92 Jan-95 Jan-98 Jan-01 Jan-04 Jan-07
Figure 5.27: Groundwater bromide concentrations at monitoring locations in the vicinity of the source site 1983 to 2008.
5.6. Contaminant Distribution 201
12
Fit Results
8
BrO3 Leachate-Soil
Equation Y = 0.07036647551 * X
Number of data points used = 11
4 Average X = 58.3195
Average Y = 4.70377
Residual sum of squares = 26.5774
Coef of determination, R-squared = 0.956988
Residual mean square, sigma-hat-sq'd = 2.65774
0
12 BROMIDE
Leachate Bromide (mg/l)
Fit Results
4 Br Leachate-Soil
Equation Y = 0.08152457005 * X
Number of data points used = 11
Average X = 45.2182
Average Y = 3.68909
Residual sum of squares = 0.779852
Coef of determination, R-squared = 0.997038
Residual mean square, sigma-hat-sq'd = 0.0779852
0
0 40 80 120 160
Soil Bromide (mg/kg)
Figure 5.28: Relationships between leachate concentration (mg l−1 ) and soil concentration (mg kg−1 )
for samples from the 2001 investigation (Atkins, 2002).
5.7. Mass of bromide and bromate at the source site 202
The partition (or distribution) coefficient, Kd , was calculated as described in Lewin et al. (1994) as
Csorbed
Kd =
Cdissolved
Where Csorbed is the sorbed contaminant concentration (mass of contaminant in mg ÷ mass of soil
in kg) and Cdissolved is the dissolved contaminant concentration (mass of contaminant in mg ÷ volume
of solution in litres).
The soil concentration (mg kg−1 ) Csoil was assumed to represent the total mass of contaminant
(sorbed plus dissolved). The dissolved contaminant concentration Cdissolved was assumed to be the
leachate concentration (mg l−1 ). Csorbed was calculated as the total soil concentration, Csoil , minus the
dissolved mass per kg of soil (Cleachate × L/S ratio2 ). Performing these calculations on the samples from
the the 2001 investigation (Atkins, 2002), gave a mean Kd of 11.8 l kg−1 for bromate and 12.7 l kg−1
for bromide.
• The site was divided up based on a 10 m × 10 m grid. Assuming a thickness of unsaturated zone
of 4 m, and a bulk density of 2000 kg m−3 , there is 8×105 kg of material in each block. The mass
of bromide in each block was then calculated as:
where MBr is the mass (in kg) of bromide in a block, CBr is the concentration (in mg kg−1 ) of
bromide in the block and Ms is the mass (in kg) of soil in block. Calculating this for each block
and summing the answers gives a total amount of bromide on site in the unsaturated zone.
1. Using just the squares where samples were taken (a total of 20 squares) gave a total amount
of 7,535.2 kg bromide on site.
2. Assuming that the squares with no analysis had concentrations the average of those squares
surrounding them, and the contamination at the boundary is zero, gave a total of 14,135 kg
bromide on site.
2 Liquid to Solid ratio. For NRA leachate analyses (Lewin et al., 1994), the Liquid to Solid ratio is 10:1 (l kg−1 )
5.7. Mass of bromide and bromate at the source site 203
• The mass of bromide removed as part of the top layer was estimated by assuming that based on
the grid pattern, the remediation proposals encompassed the excavation of a total of 17 blocks
to a depth of 0.75 m, and 23.5 blocks to a depth of 1.5 m. Calculating the mass of bromide for
each block and summing the results, gave a total of 2406 kg (566 kg + 1840 kg). Subtracting this
amount from the estimates of mass above gave a residue of 5129 kg and 11729 kg of bromide
present in the unsaturated zone after contaminated soil removal.
• Based on the available samples in the vicinity of the source zone (i.e. from Boreholes 221 and 219),
bromate and bromide is present (above background concentrations) in porewater to a maximum
depth of approximately +77.5 m OD. The maximum depth of sampling was at +75 m OD in
Borehole 223, and samples indicated very low concentrations (below or close to the MDL). It is
therefore assumed that the depth of contamination extends to between approximately +75.5 m OD
and +77.5 m OD beneath the source site.
• Ground level in the vicinity of the source zone is approximately +83.5 m OD.
• Water levels in the source zone range from +81.5 m OD to +79.0 m OD.
• Therefore, the thickness of unsaturated zone ranges from 2.0 m to 4.5 m thick and the thickness
of saturated zone ranges from 1.5 m to 6.0 m thick. The mean unsaturated zone thickness (based
on 2001-2008 data) is 4.1 m, and the mean saturated zone thickness is 3.4 m.
5.7. Mass of bromide and bromate at the source site 204
0 mg/kg 210420
083 083
210400
517020 517040 517060 517080 517100 517120 517140
Figure 5.29: Bromate soil concentration contours for 1.0 m thick grid slices based on investigation data
from 2000 and 2001 (Komex, 2000; Atkins, 2002). Estimates for total mass in the unsaturated and
saturated zones refer to minimum, mean and maximum thicknesses defined in Figure 5.31.
5.7. Mass of bromide and bromate at the source site 205
0 mg/kg 210420
083 083
210400
517020 517040 517060 517080 517100 517120 517140
+81.6 to +80.6 mOD
+80.6 to +79.6 mOD +79.6 to +78.6 mOD
Figure 5.30: Bromide soil concentration contours for 1.0 m thick grid slices based on investigation data
from 2000 and 2001 (Komex, 2000; Atkins, 2002). Estimates for total mass in the unsaturated and
saturated zones refer to minimum, mean and maximum thicknesses defined in Figure 5.31.
5.7. Mass of bromide and bromate at the source site 206
Ground Level
4.5 m
Max Water Level
+81.5 mOD
Min. ‘plume’
+79.0 mOD
Max. ‘plume’
6.0 m
1.5 m
1.5 m
6.0 m
Max recorded
contamination
+77.5 mOD
Max depth of
sampling
+75.5 mOD
Figure 5.31: Minimum and maximum saturated and unsaturated zone thicknesses.
5.8. Mass flux of bromate in groundwater migrating from the source site 207
Table 5.3: Summary of mass estimates. Estimates for total mass in the unsaturated and saturated zones
refer to minimum, mean and maximum thicknesses defined in Figure 5.31.
Z x=B
F = Dvne C dx (5.1)
x=A
where D is the depth of the contaminated ‘plume’, v is the groundwater velocity and ne is the effective
porosity
The linear velocity, v, is related to the darcy velocity, q, by q = vne , so equation 5.1 can be written
as:
Z x=B
F = Dq C dx (5.2)
x=A
R x=B
The integral x=A
C dx is the area under a graph of concentration C against distance x along the
section line from A to B (Figure 5.32).
As described in Section 5.7.2, the bromate contamination is estimated to extend to depths of be-
tween +75.5 m OD and +77.5 m OD beneath the source site. Water levels in the source zone range from
+81.5 m OD to +79.0 m OD. Therefore, if the depth of the contaminated ‘plume’ is taken as the distance
between the top of the water column and the estimated deepest recored contaminated porewater sample,
then the thickness of of ‘plume’ could range from 1.5 m to 6.0 m (Figure 5.31).
Using a range of 1.5 m to 6.0 m for D, and a darcy velocity q of 0.9 m d−1 based on the results
of the single borehole dilution tests (Section 3.6.1) at the nearby Harefield House (location 226) and
Nashe’s Farm (location 019) boreholes, groundwater flux estimates for bromate (Figure 5.32 range from
5.9. Previous representations of the ‘Source Term’ 208
6.3 kg d−1 to 41 kg d−1 (572 to 3770 kg y−1 ) for a ‘thick plume’ and 1.6 kg d−1 to 10 kg d−1 (2289
to 15080 kg y−1 ) for a ‘thin plume’. Flux estimates for bromate apparently decrease with time between
2001 and 2003.
These flux estimates seem extremely high in relation to the amount of bromate present at source
site in the unsaturated and saturated zone (Section 5.7.2). This suggests that the dominant contributor to
the groundwater concentrations is bromide/bromate in the saturated zone porewater, and that the mass
of bromide/bromate in the porewater has not been fully accounted for in the mass calculations. Also, the
‘plume’ may extend deeper than accounted for in the mass calculations and/or concentrations may be
higher in locations where samples were not taken.
Bromate Bromate
160000 Apr-Jun 2002 80000 Apr-Jun 2003
Area under graph 6.0 m thick plume
(ug/l * m) Flux (kg/d) Flux (kg/y)
120000 60000
Oct-Dec 2001 7645509 41.3 15080
Jan-Mar 2002 3253520 17.6 6417
Apr-Jun 2002 2863282 15.5 5647 80000 40000
Jul-Sep 2002 2404950 13.0 4743
Oct-Dec 2002 2143395 11.6 4228 40000 20000
Jan-Mar 2003 1263853 6.8 2493
Apr-Jun 2003 1735090 9.4 3422 0
Bromate Bromate
120000 Jul-Sep 2002 100000 Jul-Sep 2003
80000
80000 60000
40000
40000
20000
0
Bromate concentration (µg/l)
Bromate
400000 Oct-Dec 2001 Bromate
5.9. Previous representations of the ‘Source Term’
Bromate
100000 Oct-Dec 2002 60000 Oct-Dec 2003
300000
80000
40000
200000 60000
0
Bromate concentration (µg/l)
0 40 80 120
Distance x along flux plane (m) 0 40 80 120 0 40 80 120
209
Figure 5.32: Estimates of bromate groundwater flux from the ‘source zone’ using equation 5.2 and the area under a concentration profile taken across a flux plane through the
R x=B
source zone. The area under a graph represents the integral x=A C dx. The flux plane is shown in Figure 5.23.
5.10. New Conceptual Models for Contaminant Release 210
concentration was assigned a range of 1.7 mg kg−1 to 273 mg kg−1 based on the analytical results of
samples collected from the unsaturated soil zone at these locations.
The probabilistic CONSIM modelling package was then applied to predict leaching of soil contam-
inants through the unsaturated zone and to predict concentrations arising in the Putty Chalk aquifer. The
subsequent dilution within the saturated Chalk aquifer (a 10 m thick mixing zone comprising saturated
Putty Chalk and Blocky Chalk) was also simulated. The results of the simulations indicated bromate con-
centrations close to observed groundwater concentrations at monitoring locations 219 and221. Atkins
(2002) concluded that this indicated that there was a ‘significant pollutant linkage’ (SOURCE → PATH-
WAY → RECEPTOR) between the bromate contamination source within the unsaturated zone soils and
groundwater within the Chalk aquifer receptor beneath the site, via a pathway of leaching through the
unsaturated zone.
A constant source term is not realistic for the site: the bromate source will be depleted as mass is
transported down-gradient and a steady concentration would not be expected to be maintained into the
future. Also, monitoring data for bromide in groundwater at the site show that concentrations have de-
creased considerably between 1984/1985 and 2000/2001 (Section 5.6.2.2). There is also some indication
that concentrations have declined at locations in the vicinity of the source site (Section 5.6.3). If bromate
contamination has followed a similar history to bromide contamination, then it would be expected that
bromate concentrations at the source site would have also shown some decline.
Based on the distribution of bromide from investigations in 1983, 1984 and 1985, it is apparent
that by the end of the operational lifetime of the chemical works, considerable bromide (and most likely
bromate), had accumulated in the unsaturated and saturated zone beneath the source area. Figure 5.34
presents a conceptual model for the release of bromate and bromide to groundwater beneath the source
zone. The mechanism by which the contamination was released and accumulated between 1955 and 1983
is unknown. The main areas of uncertainty in the history of bromate and bromide release to groundwater
beneath the site are summarised in Table 5.4.
5.10. New Conceptual Models for Contaminant Release 211
Table 5.4: Main areas of uncertainty in the history of bromide and bromate release to groundwater
beneath the source site.
Form of input
The bromide concentrations in soil porewater and
groundwater recorded at the source site between 1983 and
The form of release could range from a
1987 are significantly higher than concentrations recorded
relatively constant input through continuous
in 2000 and 2001. The buildings were cleared and the site
leak/discharge, to an input over a short
was left to free-drain in 1984. The site clearance is likely to
period of time as a result of a catastrophic
have coincided with a substantial increase in infiltration
leak/discharge or a recharge pulse.
rate, and therefore a pulse of recharge. This recharge
pulse may have been important in transporting bromide
Release may have occurred as a focused
(and bromate) from unsaturated zone to the saturated
input over a small area of the site (e.g. via
zone.
sumps) or may have occurred as a more
diffuse input over a larger area.
Similarity of bromide and bromate release Based on the site investigation data for 2000 and 2001, the
history spatial distribution of elevated bromide and bromate in
porewater and groundwater shows similarities. Also,
The release history of bromate and bromide bromate and bromide concentrations are positively
may or may not be related and show correlated based on groundwater monitoring data from
similarities in the form and timing of bromate 2000 to 2007.
input.
However, the lack of data for bromate concentrations prior
If the predominant mobilisation of to 2000 means that it is uncertain whether or not bromate
contamination occurred via a pulsed concentrations were similarly elevated in the early 1980s.
recharge event, the form of bromide and The only concentrations of bromate available are for
bromate release from 1984 is likely to be shallow soil in 1983, and all samples were below detection
similar. limits. Therefore, it is possible that bromate contamination
release occurred later than bromide contamination (hence
However, if releases prior to 1983 have been bromate not as widespread in 1983), and/or at lower
more important in the mobilisation of bromide relative quantities.
to groundwater, then due to lateral
separation of bromide and bromate It is also possible that bromate was elevated in 1983 but
production areas, the relative magnitudes was not detected because release mechanism meant that
and timing of release may differ its distribution was deeper than 1.5 m (e.g. a more focused
considerably. input into drainage via sumps).
5.10. New Conceptual Models for Contaminant Release 212
Figure 5.33: The combined ‘source zone’ (centre figure) based on the locations of high concentrations
of bromate (left hand figure) and bromide (right hand figure) in groundwater
Figure 5.34: Conceptual Model for bromate and bromide release from the source zone.
5.10. New Conceptual Models for Contaminant Release 214
operation of the factory, much of the site was covered with buildings or hard-surfaces which would have
restricted the amount of recharge. However, a much higher rate of recharge would have been possible
when the site was cleared and left to free-drain (end of 1983 to beginning of 1987). Leaching would have
the effect of transporting bromate and bromide ions through the unsaturated zone to groundwater within
the saturated putty chalk. The large decrease in bromide concentrations encountered in 2000 and 2001
compared to those recorded between 1983 and 1985 in comparable locations and depths, suggests that
high recharge rates during 1984 to 1987 may have played an important role in leaching out contamination
from the unsaturated zone profile.
Assuming recharge to occur via ‘piston flow’ (downward movement of water that has infiltrated at
the surface occurs via vertical drainage through unsaturated matrix porewater), a minimum time for the
complete leaching of the contaminant from the soil profile can be estimated by calculating the length of
time it would take to replace the water contained within the unsaturated zone soil profile:
VU
Time for replacement of water in unsat. zone = VR
Where
Following the estimates of Roberts (2001), assuming a 4 m unsaturated zone with moisture content
ranging from 11 % to 13 % and recharge rate 150 mm y−1 to 350 mm y−1 , the minimum time for
replacement is between 3.6 years and 1.3 years, with average values of approximately 2 years. These
estimates predict that all the water present in the unsaturated zone could be replaced over a period of
approximately 2 years, which implies that if the site was cleared at the end of 1983, and development
took place at the beginning of 1987, all the water present in the unsaturated zone at the time of site
clearance would have been replaced by infiltrated rainwater. Therefore, if the bromate and bromide ions
behave conservatively (i.e. they are transported at the same rate as the water in which they are dissolved),
then the ‘free’ bromide and bromate would have been almost completely removed from the unsaturated
zone over this period.
However, it is clear that although bromate concentrations are generally low in the upper part of
the unsaturated zone, relatively high concentrations still remain in the lower unsaturated zone (SUZ).
The presence of discontinuous clay horizons within the fluvial-glacial deposits may have the effect of
restricting recharge in some areas whilst concentrating recharge through other ‘windows’ in the clay.
Where clay horizons restrict recharge, this may account for the continued presence of bromide and
bromate within the unsaturated zone (present either as a result of transport from above, or introduction
during high groundwater levels), despite high potential rates of leaching.
The bromide ion is generally thought to be conservative (Section 1.7) and so is transported at the
same rate as the water in which it is dissolved. If the bromate ion behaves like the bromide ion, i.e.
conservatively, then bromate will also be transported at the same rate as the water in which it is dissolved.
However, the results of leachate tests carried out in 2001 (Section 5.6.5) at the site indicates that some
5.10. New Conceptual Models for Contaminant Release 215
partitoning of bromate and bromide occurs between the soil phase and dissolved phase. The downward
movement of bromate and bromide is therefore inhibited by soil interaction, and the time needed for
complete removal from the soil is likely to be somewhat increased compared to the movement of soil
porewater.
5.10.1.4 Diffusive exchange with immobile porewater within the saturated zone
The (blocky) Chalk is considered to behave as a double-porosity aquifer with groundwater flow occur-
ring predominantly through fissures, and the high porosity, low permeability matrix providing storage
(Section 2.5). Diffusive exchange between the mobile fissure water and the immobile matrix water will
occur during groundwater flow (Section 2.9.4).
Assuming initially contaminated fracture water and uncontaminated matrix water, bromate diffu-
sion into the matrix water would have acted to retard the transport of bromate down-gradient. At a later
stage, which may or may not have been reached at the SLC site, when the original source of contam-
ination within the fracture water has ceased, the direction of diffusion will be from the contaminated
3 Estimates of recharge (the portion of rainfall that percolates down to the water table) for free draining site are between
150 mm y−1 and 350 mm y−1 compared to approximately 60 mm y−1 for the site with predominant cover of hardstanding.
5.11. Source terms for bromide and bromate release from the source site 216
matrix water to the less contaminated fracture water, which will act to prolong the period of bromate
contamination.
5.10.1.5 Seasonal mobilisation of porewater within the zone of water table fluctuation
Within the zone of water table fluctuation, also referred to as the seasonally unsaturated zone (SUZ),
fractures are periodically filled and drained when the water table fluctuates. Due to the low permeabil-
ity of the Chalk matrix, the matrix remains saturated. Diffusive exchange between fracture water and
matrix water occurs when the fractures are saturated. This may provide a seasonal source (or sink) to
groundwater in the saturated putty chalk.
This process may explain the concentration of bromate and bromide within the SUZ. During periods
of high groundwater levels, diffusion is likely to have occurred between contaminated groundwater and
uncontaminated porewater in the SUZ. Periodic repetition of this process results in the accumulation
of contamination within porewater of the SUZ. At a later stage of groundwater plume evolution, or if
initially high concentrations are present in porewater (e.g. from leaching of contamination from above),
diffusive exchange between contaminated porewater and less contaminated groundwater would occur
seasonally, proving a seasonal source of bromate to groundwater.
5.11 Source terms for bromide and bromate release from the
source site
In order to encompass the range of possible scenarios for bromate release to groundwater beneath the
source site, three source term scenarios are developed within this chapter. They are summmarised below:
• Scenario A - ‘Catastrophic Release’ This scenario envisages a sudden, large, release of bromide
and bromate to the unsaturated zone at some point in the operational history of the factory.
• Scenario B - ‘Steady Seepage’ This scenario envisages a steady discharge of bromide and bro-
mate to the unsaturated zone during the operational lifetime of the factory.
• Scenario C - ‘Direct Release’ This scenario envisages a direct release of bromate to the saturated
zone (by-passing the unsaturated zone) for a period of time over the operational history of the
factory.
The three conceptual scenarios for bromate input to, and release from, the ‘source zone’ are quan-
tified in the following sections, using estimates of bromide and bromate mass in the unsaturated and
saturated zone of the source site (Section 5.7.2) to constrain the range of possible source histories.
concentration, CL , in mg l−1 ):
CS
Kd = (5.3)
CL
At time t, the total mass in the unsaturated zone, M (t), is represented by the equation:
S−R
M (t) = M0 exp−Kt + (1 − exp−Kt ) (5.4)
K
Where S is the rate of ‘seepage’ of mass to the unsaturated zone, R is the rate of mass removal from
the unsaturated zone (by excavation, remediation etc.), both of which are assumed to be constant over
defined time periods, and K is a constant defined as:
1000r
K= (5.5)
DρKd
where r is the infiltration rate or recharge rate, D is the thickness of the unsaturated zone, ρ is the bulk
density of soil within the unsaturated zone, and Kd is the distribution coefficient.
The rate of leaching of contaminant from the unsaturated zone, L(t), is represented as:
The total mass of contaminant leached from the unsaturated zone between time T1 and T2 is then
given by:
Z T2
ML (t) = L(t)dt (5.7)
T1
L(t)
CL (t) = (5.8)
1000rAθm
where θm is the mobile porosity (the effective porosity) and A is the site area open to infiltration.
5.11.2 Constraints
The source term in bounded by the following constraints:
• Condition 1 – the mass in the unsaturated zone at the end of 1983 should correspond to the
observed mass estimate for the unsaturated zone in 1984/1985;
• Condition 2 – the total mass input up until 1983 should be at least the observed mass estimate for
the saturated zone in 1984/1985;
• Condition 3 – the mass in the unsaturated zone in 2000/2001 should correspond to the observed
mass estimate for the unsaturated zone in 2000/2001;
• Condition 4 – the total mass input up until 2008 should be at least the amount of mass estimated
to have been removed by Hatfield and other abstractions between 1981 and 2008.
5.11. Source terms for bromide and bromate release from the source site 218
Figure 5.35: Derivation of equations for mass of bromide/bromate in the unsaturated zone and the rate
of input of bromide/bromate from the unsaturated zone to the saturated zone.
5.11. Source terms for bromide and bromate release from the source site 219
Bromate conc. (mg kg−1 ) = 0.36 × Bromide conc. (mg kg−1 ) + 32.1 (mg kg−1 )
In order to fit the observed bromate mass results, a value of Kd of 0.23 l kg−1 is required (Fig-
ure 5.36). Therefore a Kd of 0.23 l kg−1 is used subsequently for calculating the source term for
bromate.
The source terms for Scenario A and Scenario B from 1984 onwards are illustrated in Figure 5.37
as bromide and bromate concentrations released to the saturated zone from 1984 onwards.
40000
Equation of line:
Bromide (kg)
M1986 = 18115 kg
Using the values for parameters r, S, R and M S = 0, R = 0 kg y-1
indicated on chart B r = 500 mm y-1
M1985 = 24907 kg
22026
Kd = 12 litres kg-1 With Kd = 12 litres kg-1
too much partitioning occurs
8103 and equations cannot be fit to
observed mass of bromide
Equation fit to observed mass Kd = 0.20 litres kg-1
2981 in 1985 and 2001by adjusting Kd
Kd = 0.20 litres kg-1 allows equations to be fit to
observed mass of bromide
Bromide (kg)
1987 < t ≤ 2001
1097 S = 0, R = 0 kg y-1 M2001 = 812 kg
r = 290 mm y-1
148
Figure 5.36: Equations for bromide mass, fit to observed values from 1985 and 2001. Parameters are defined in Figure 5.11.1
2
1x10
Scenario A and Scenario B 1984+ Bromide conc.
Bromate conc.
60
1
1x10
0
1x10
50
-1
1x10
40 -2
1x10
1x10-4
1980 1990 2000 2010 2020 2030 2040 2050 2060
10
5.11. Source terms for bromide and bromate release from the source site
0
1980 1985 1990 1995 2000 2005 2010 2015 2020
221
Figure 5.37: Bromide and bromate concentrations for Scenario A and Scenario B from 1984 into the future.
5.11. Source terms for bromide and bromate release from the source site 222
taken as 11,140,000 kg of bromide and 221,600 kg of bromate input over the year of 1965. Table 5.5
summarises how the source term relates to the four conditions in Section 5.11.2. Scenario A is illustrated
as a source term in Figure 5.40.
Table 5.5: Mass predicted by source history scenarios A and B compared to observed mass constraints.
Condition 4 is based on an estimate by Buckle (2002) of the mass removed at Hatfield and Essendon
between 1981 and 2000.
Bromide - SCENARIO A
10000000 10000000
1000000 1000000
100000 100000
10000 10000
1000 1000
100 100
10 10
1 1
1955 1960 1965 1970 1975 1980 1985 1990 1995 2000 2005 2010 2015 2020
Bromide - SCENARIO B
40000 4000
30000 3000
20000 2000
10000 1000
0 0
1955 1960 1965 1970 1975 1980 1985 1990 1995 2000 2005 2010 2015 2020
Max Br unsat
Mean Br unsat
Min Br unsat
Rate of mass leaching from unsaturated zone (kg yr-1)
Mass in the unsaturated zone (kg)
Bromate - SCENARIO A
1000000 1000000
100000 100000
1000 1000
100 100
10 10
1 1
1955 1960 1965 1970 1975 1980 1985 1990 1995 2000 2005 2010 2015 2020
Bromate - SCENARIO B
16000 16000
12000 12000
8000 8000
4000 4000
0 0
1955 1960 1965 1970 1975 1980 1985 1990 1995 2000 2005 2010 2015 2020
Max Br unsat
Mean Br unsat
Min Br unsat
Rate of mass leaching from unsaturated zone
Mass in unsaturated zone
Bromide concentration
2
2.0x10
1.9x102
2
1.8x10
1.7x10
2 Scenario A
1.6x10 2 Scenario B
1.5x102
1.4x102
Bromide concentration (mg litre-1)
2
1.3x10
2
1.2x10
2
1.1x10
1.0x102
9.0x101
8.0x101
7.0x101
1
6.0x10
5.0x101
4.0x101
3.0x101
2.0x101
1.0x101
0.0x100
Bromate concentration
3.5x101
1
3.0x10
2.5x101
Bromate concentration (mg litre-1)
2.0x101
1.5x101
1
1.0x10
5.0x100
0
0.0x10
Figure 5.40: Bromide and bromate concentrations for Scenario A and Scenario B between 1955 and
1984. After 1984 concentrations proceed as in Figure 5.37.
5.12. Verifying source terms with observed down-gradient concentrations 226
2000 300
200
1000
100
0
1950 1970 1990 2010 1950 1970 1990 2010 2030 2050
500
3000
Bromate concentration (mg/l)
400
Bromate concentration (mg/l)
2000 300
200
1000
100
0
1950 1970 1990 2010 1950 1970 1990 2010 2030 2050
500
3000
Bromate concentration (mg/l)
400
Bromate concentration (mg/l)
2000 300
200
1000
100
0
1950 1970 1990 2010
1950 1970 1990 2010 2030 2050
implanted in Excel. The spreadsheet accesses the code, held in a dynamic link library, via Visual Basic.
In order to evaluate the potential for the DP1-D code was used to simulate bromide and bromate
concentrations down-gradient along a one-dimensional flow line from the source zone. The simulated
results were compared to observed concentrations at three locations with bromide monitoring data avail-
able for 1983-1987 and 2000-2008, and bromate data for 2000-2008 only.
The parameters selected (Figure 5.42; Sections 5.12.1 to 5.12.6) were chosen as best available esti-
mates from investigation data (i.e. the model was not calibrated to produce the observed concentrations.)
The conceptual model for bromate and bromide release to groundwater beneath the site considers
the ‘putty chalk’ to represent a low permeability layer between the base of the unsaturated zone, and
groundwater migrating off-site in the ‘blocky chalk’. Groundwater flow is therefore assumed to be es-
sentially vertical within the putty chalk layer. Groundwater flow is simulated vertically through the putty
chalk layer to the saturated blocky chalk below the site, and then groundwater is simulated horizontally
along a groundwater flowline (Figure 5.42).
ρga2
K=
12µ
where ρ is the fluid density, µ is the fluid viscosity, and g is the acceleration due to gravity.
2bΦ
θim = (5.10)
2b + a
Therefore, the immobile porosity θim is estimated as 0.38 for the putty chalk and 0.38 for the blocky
chalk.
Simulated bromate concentrations groundwater show good agreement with the observed bromate
concentrations at monitoring location 225 and 028 (Figure 5.44), but simulated concentrations at loca-
tion 019 are at the lower limit of the the observed concentrations.
5.12.7.2 Scenario B
For Scenrio B, simulated bromide concentrations (Figure 5.45) and bromate concentrations (Figure 5.46)
are generally lower than observed concentrations by approximately one order of magnitude.
5.12.7.3 Scenario C
Simulated bromate concentrations for Scenario C (Figure 5.47) show good agreement with the observed
bromide concentrations at all three of the monitoring locations.
BH2-[84] ; BH3-[84]
1000 Observed
BH 221; BH 219
Observed
Monitoring
Source Term - Scenario A
100 Location
028 Concentrations simulated along
a pathline from the source zone
10 to locations 150 m, 850 m and
1200 m down hydraulic gradient
1 Source
1000 Location 225 / B1 8 Location 028 - Orchard Garage BH 8 Location 019 - Nashe's Farm BH
Obs. Groundwater
Obs. Porewater
Sim. Groundwater
800 Sim. Porewater
6 6
5.13. Summary and conclusions
600
4 4
400
2 2
200
Figure 5.43: Comparison of simulated bromide concentrations for source history Scenario A and observed concentrations at three monitoring locations.
SOURCE TERM
10000 input to the saturated Putty Chalk
BH 221; BH 219
1000 Observed
Source Term - Scenario A
Monitoring
100 Location
028 Concentrations simulated along
a pathline from the source zone
10 to locations 150 m, 850 m and
1200 m down hydraulic gradient
1 Source
6 Location 225 5 Location 028 - Orchard Garage BH 5 Location 019 - Nashe's Farm BH
Sim. Groundwater
Sim. Porewater
Obs. Groundwater
Obs. Porewater 4 4
5.13. Summary and conclusions
4
3 3
2 2
2
1 1
Figure 5.44: Comparison of simulated bromate concentrations for source history Scenario A and observed concentrations at three monitoring locations.
SOURCE TERM
104 input to the saturated Putty Chalk
3 BH2-[84] ; BH3-[84]
10 Observed
BH 221; BH 219
2 Observed
10 Monitoring
Source Term - Scenario B
Location
1
028 Concentrations simulated along
10 a pathline from the source zone
to locations 150 m, 850 m and
1200 m down hydraulic gradient
100
Source
-1
1000 Location 225 / B1 10 Location 028 - Orchard Garage BH 10 Location 019 - Nashe's Farm BH
Obs. Groundwater
Obs. Porewater
Sim. Groundwater
800 Sim. Porewater 8 8
5.13. Summary and conclusions
600 6 6
400 4 4
200 2 2
Figure 5.45: Comparison of simulated bromide concentrations for source history Scenario B and observed concentrations at three monitoring locations.
SOURCE TERM
103 input to the saturated Putty Chalk
BH 221; BH 219
102 Observed
Source Term - Scenario B
Monitoring
101 Location
028 Concentrations simulated along
a pathline from the source zone
100 to locations 150 m, 850 m and
1200 m down hydraulic gradient
-1 Source
10
6 Location 225 5 Location 028 - Orchard Garage BH 5 Location 019 - Nashe's Farm BH
Sim. Groundwater
Sim. Porewater
Obs. Groundwater
Obs. Porewater 4 4
5.13. Summary and conclusions
4
3 3
2 2
2
1 1
Figure 5.46: Comparison of simulated bromate concentrations for source history Scenario B and observed concentrations at three monitoring locations.
SOURCE TERM
3000 input to the saturated Putty Chalk
2000 Monitoring
Location
028 Concentrations simulated along
a pathline from the source zone
to locations 150 m, 850 m and
1200 m down hydraulic gradient
1000
Source
10 Location 225 5 Location 028 - Orchard Garage BH 5 Location 019 - Nashe's Farm BH
Sim. Groundwater
Sim. Porewater
Obs. Groundwater
8 Obs. Porewater 4 4
5.13. Summary and conclusions
6 3 3
4 2 2
2 1 1
Figure 5.47: Comparison of simulated bromate concentrations for source history Scenario C and observed concentrations at three monitoring locations.
5.13. Summary and conclusions 237
Figure 5.48: Concurrent matrix and fissure concentrations are required to determine at which point along
the concentration-time graph a particular fissure concentration represents.
contaminant input to the saturated zone beneath the site is the leaching of bromide and bromate mass
from the unsaturated zone. The scenarios are constrained by estimates of the observed mass of bromide
and bromate in 1985 and 2001. Scenario C (‘direct release’) assumes that bromate is released direct to the
saturated zone, by-passing the unsaturated zone. Scenario C is constrained by the observed groundwater
concentrations at the source site in 2001.
The one-dimensional double-porosity transport code, DP1-D, has been used to simulate concentra-
tions in groundwater down-gradient of the source site. Simulated concentrations using the source term
scenarios A and C show relatively good agreement with observed groundwater concentrations at loca-
tions 150 m, 500 m and 1000 m down-gradient of the source site. However, the relatively short period of
time for which monitoring data are available, combined with the large seasonal variations in concentra-
tions, means that the trends are difficult to discern, and robust conclusions cannot be made as to whether
or not the simulations are representative of the observed data.
238
Chapter 6
• To review the previous modelling approaches that have been implemented for bromate and/or
bromide contamination in Hertfordshire;
• To use this model to provide predictions for the likely long-term evolution of bromate concentra-
tions at key output locations.
source term which have been discussed in Section 5.9, and indicate that the mass input to the aquifer is
likely to have been considerably underestimated by the Chemfix modelling.
The model parameters were derived from literature data, and site-specific data where available,
and are reasonably consistent with the likely parameters for the Chalk aquifer in the region reviewed in
Chapter 2. A constant source term was assumed from 1970 to 2000 of 1000 µg l−1 at Orchard Garage
and 100 µg l−1 at Hatfield. Between 2000 and 2004, monitoring data at Orchard Garage and Hatfield
respectively were used to represent the source terms. From 2004 into the future constant source terms of
1000 µg l−1 at Orchard Garage and 244 µg l−1 at Hatfield were assumed.
The simulated bromate concentrations at Hatfield P.S. and Essendon P.S. corresponded well to the
‘average’ observed bromate concentration trends. Simulated concentrations at Hoddesdon corresponded
reasonably well to the ‘average’ observed bromate concentration trends. The simulations did not cap-
ture the seasonal variation evident in observed bromate concentrations due to the non-seasonality of the
model, the relatively constant source-term, and the smoothing effect of the double-porosity diffusive ex-
change. All scenarios indicated increasing concentrations over time. This rising trend was interpreted by
Atkins (2004) as suggesting that the ‘plume’ is not in steady state, i.e. fracture and matrix concentrations
had not yet reached equilibrium at any of these locations.
significantly higher than would be expected purely from the effects of dispersion, with the results for
Turnford and Amwell Hill in particular suggesting that a further mechanism of contaminant transport,
such as separate direct flow pathways leading to these wells, was important.
the models (Buckle, 2002), the representation of the karst system within the model is limited, probably
due to deficiencies in the conceptual model and inadequate data to properly parameterise the system at
that time. Cook (2010) used his new conceptualisation of the Hertfordshire karst flow system, along
with hydrodynamic parameters determined by catchment-scale tracer testing, and the single borehole
dilution tests described in this thesis (Section 3.6.1), to develop a suite of new models incorporating
representations of the karst flow system.
Initially, Cook (2010) developed a steady-state subset MODFLOW model of the NNR model to
allow faster execution times, simpler initial calibration to long-term heads and flows and improved model
stability. The 200 m by 200 m grid of the NNR model was retained. The model uses an EPM approach,
and incorporates a karst zone along palaeogene boundary which is represented as a zone of high hydraulic
conductivities, between one and a few cells wide. The karst zone also extends westwards into the Vale of
St. Albans. The model was used to simulate flowlines (using MODPATH) for the 2008 bacteriophage
tracer tests, and parameters calibrated to achieve good representation of all three tracer breakthroughs
based on the advective transport routes indicated by MODPATH flowlines.
Cook (2010) then converted the calibrated steady-state flow and transport model to a transient flow
and transport model. The transient flow model was found, for the majority of locations, to replicate the
magnitude of head variation and also the seasonal behaviour and trends relatively well. The transport
model (using MT3D-MS) was run for three source scenarios (see discussion of source terms in Sec-
tion 6.2.5.1), as well as constant concentration source of 5000 µg l−1 for comparison with the NNR
model of Atkins (2005). Cook (2010) found that the simulations showed closer agreement with travel
time of the recharge pulse mass input and observed data at receptors if the westward extension of the
karst EPM into the Vale of St Albans was removed. With karst in the Vale of St. Albans, weakly pos-
tulated on the basis of the tracer tests, the timing of simulated peak concentrations precedes that of all
observations. However, with the karst zone removed, simulated concentrations are lower, and travel
times slower, due to additional dispersion and double-porosity attenuation.
Simulations using the constant concentration source of 5000 µg l−1 were able to closely replicate
observed bromate concentrations to Hatfield in magnitude and spatial distribution. Simulations using
the previous estimates (now superseded) of the three source term scenarios (Section 6.2.5.1) resulted in
concentrations that were universally lower than observed. Cook (2010) concluded that these source terms
were too low and increased them by a factor of three. The revised source term Scenario B was found
to result in relatively good spatial and temporal representation of bromate concentrations compared to
observed data. The situation as modelled by the revised source term of Cook (2010) suggests that current
levels of bromate in the aquifer are the result of the passage of the 1983-1987 high concentration recharge
pulse from the source zone.
The model produces a relatively stable ‘plume’ in the Vale of St Albans area after approximately 10
years of bromate input. Breakthrough to the Lee Valley occurs shortly afterwards due to the rapid trans-
port within the karst system, and is strongly seasonal. The model predicts that bromate concentrations
in excess of the drinking water standards persist within the Vale of St Albans for the modelling period
6.2. Previous modelling approaches for Bromide and Bromate in the Chalk 242
(i.e. at least up until 2050). For locations in the Lea Valley, dilution in the karst system acts to reduce
concentrations significantly to less than drinking water standards once the high mass flux associated with
the 1983-1987 recharge pulse has declined. However periodic seasonal pulses of a around 1 µg l−1 oc-
cur at locations in the Lee Valley for the remainder of the modelling period. These pulses are strongly
influenced by karstic seasonal dilution.
Cook (2010) lists the main areas of uncertainty that limit the effectiveness and confidence in the
MODFLOW/MT3DMS predictive modelling using the currently available data:
• The description of the source term and its implementation in the model;
• A limited period of observation data in relation to the likely duration of the bromate contamination
and with respect to the timing of the breakthroughs;
• Uncertainty with respect to the extent to which the chalk matrix has become contaminated and
detailed parameterisation of that process;
• Deficiencies in the model representation of both dual porosity and karstic transport and the trans-
fers between them.
∂Cm ∂Cm ∂ 2 Cm
θm = −q + θm Dim − ζ (Cm − Cim ) (6.1)
∂t ∂x ∂x2
6.2. Previous modelling approaches for Bromide and Bromate in the Chalk 243
Bromate - SCENARIO A
1000000
100000
10000
Bromate (kg)
1000
100
10
1955 1960 1965 1970 1975 1980 1985 1990 1995 2000 2005 2010 2015 2020
Bromate - SCENARIO B
16000
12000
Bromate (kg)
8000
4000
1955 1960 1965 1970 1975 1980 1985 1990 1995 2000 2005 2010 2015 2020
Max Br unsat
Mean Br unsat
Min Br unsat
Rate of mass leaching from unsaturated zone
Mass in unsaturated zone
Superceeded-input rate
Superceeded-mass
1984-1985
Superceeded-input ratex3
Figure 6.1: Comparison of the superseded versions of the the source terms used by Cook (2010) to the
current versions in this thesis.
6.2. Previous modelling approaches for Bromide and Bromate in the Chalk 244
∂Cim
θim = ζ (Cm − Cim ) (6.2)
∂t
Where Cm and Cim are the concentrations in the mobile and immobile domains respectively, θm
and θim are the porosity of the mobile and immobile domain respectively, q is the darcy flux, and ζ is
the first-order mass transfer coefficient.
The first order mass transfer coefficient approach is only an approximation to the diffusion pro-
cess, which is more accurately represented by a Fickian approach (as used in DP1D and MAP codes).
The predictions made by the mass transfer approach and Fickian diffusion converge as a steady-state
condition between fracture and matrix porewater solute concentration is attained (Barker, 1985b). At
early times the MT3D-MS approach will over-estimate solute concentration in fracture water by un-
derestimating the diffusive flux into the immobile matrix porewater domain. The rate of transfer from
the mobile to immobile domain is too slow as it does not account for the infinitely steep concentration
gradient that initially exists at the mobile/immobile interface. The mass transfer approach will therefore
predict a solute ‘plume’ to have moved further than that modelled using the Fickian exchange approach.
When the concentration gradient is reversed and the the movement of solute is from the immobile to
the mobile domain, the reverse will be the case such that the rate of transfer of solute back into the mobile
groundwater in the fracture system predicted by a first order mass transfer coefficient will be slower
than predicted by a Fickian approach. Consequently the mass transfer approach will underestimate the
fracture water solute concentrations.
Compounding this deficiency of a mass transfer approach, the MT3D-MS code of Zheng and Wang
(1999) does not allow back-diffusion from the immobile matrix domain to the mobile fissure water do-
main. Therefore the persistence of contamination will be considerably underestimated as the secondary
source of contaminant is effectively ‘lost’ into a matrix sink.
An indication of whether the system is likely to approach a steady-state within the time frame of
interest can be given by calculating the time for diffusion across a matrix block (Barker, 1993). If the
system is likely to achieve a steady state within the time frame for which predictions are required then
it is reasonable to model it using a mass transfer coefficient approach. However, if accurate predictions
are required for early or late times, a Fickian approach is recommended. Based on typical parameters for
the chalk west in the Vale of St Albans, tcb = 2.89 × 104 days ≈ 79 years (Appendix F). For the karst
system east of Hatfield, tcb ranges from 9 × 104 to 4 × 105 days, or approximately 250 to 1100 years
(Appendix F). Therefore, the mass transfer approach is not likely to be valid for bromate in Hertfordshire.
Furthermore, as noted by Cook (2010), an equivalent porous media model cannot represent simul-
taneously the conduit and fracture systems: model cells must either be an EPM of the karst conduits or
an EPM of the non-karstic fissured aquifer system. The best replication of Vale of St Albans transport in
MODFLOW and MT3D-MS is achieved using a non-karst double-porosity approximation. Conversely,
the majority of transport east of Hatfield is represented by an EPM representing karst transport. The role
of double-porosity diffusive exchange within the fracture system east of Hatfield is not simulated.
6.3. Development of a Multiple Analytical Pathways Approach 245
Figure 6.2: Conceptual and mathematical basis for the Multiple Analytical Pathways of Barker (2001).
6.3. Development of a Multiple Analytical Pathways Approach 247
Mathematical basis
A clear advantage is that only concentrations at points and times of interest need be
computed. E.g. times can be increased on a logarithmic scale when studying long-term
behaviour of dual-porosity systems.
Figure 6.3: Conceptual and mathematical basis for the Multiple Analytical Pathways of Barker (2001).
6.3. Development of a Multiple Analytical Pathways Approach 248
computationally fast. A disadvantage of the methodology is than no account can be made of variations
in abstraction volumes, because this is a steady-state model that specifies equilibrium flow pathways
from the water table to the abstractions. The necessary assumption that the transport equations are linear
restrict the type of systems that can be modelled. For example, when sorption is important, only the
linear isotherm is amenable to simulation with MAP.
The MAP model used by Williams et al. (2003) did not permit branching or joining streamtubes
(although in principle, branching or joining networks of streamtubes are possible provided each junction
is characterised by a simple additive relation between the streamtube concentrations and fluxes). In
addition, the model was not coded to give matrix concentrations as an output.
in GoldSim. However, the immobile (matrix) concentrations simulated by the ‘diffusion cells’ approxi-
mation in GoldSim did not compare well to the average immobile (matrix) concentrations simulated for
a similar scenario in DP1D.
The conclusion from this exercise was that neither MAP or GoldSim were currently appropriate
for the aims of the modelling for Hertfordshire bromate contamination as they are unable to provide
predictions for matrix porewater concentrations, which could be used in the future to validate the model
predictions for long-term fissure water concentrations.
• the geometry of the matrix blocks (currently only a slab geometry is modelled by the code);
• cm - the average solute concentration in the matrix at the downstream end of a branch;
• ctav - the porosity weighted average concentration for fractures and matrix.
Where branches converge at a node, the resulting solute concentration, cn , is given by the sum
of the concentrations, cf , at the end of each of these branches weighted according to the volumetric
groundwater fluxes through each.
v e r Rib
rL R
ea
Double-porosity pathway
River Beane
(o
Conduits
rL
ee
)
Ne
w
Ri v
!
. Spring
!
. !
.
Harefield
!
.
House Comet Way Lynchmill
Hatfield
Quarry
!
. Spring
!
.
!
.
s
an
r
lb
Ri v e
tA
S
Ne w
of
le
Va
e !Water
.
ln
Co End
r
Ri v e
6.5. A Network Model for Hertfordshire
Riv e r L ea
rne
ou
B
e
rin
Cathe
Mymm shal l Brook
0 0.5 1 2 3 4 5 © Crown Copyright/database right 2008. An Ordnance Survey/EDINA supplied service.
Kilometers Geological Map Data © NERC 2008.
Figure 6.4: Nodes and branches represented in the Network Model for Hertfordshire. Note that branches are shown schematically as straight-line connectors and are not
251
6.5.2.3 Sandridge to Comet Way, via Harefield House and Hatfield Quarry
For Sandridge to Hatfield, the parameters were estimated from a combination of typical values from the
literature and values obtained from fieldwork in the area.
The parameters a (fracture aperture) and b (half block thickness) determine the value of σ, and along
with Dim , the value of tcb (sections 5.12.1 to 5.12.3). Based on the review of data from the literature
(Section 2.7.3.3), a representative fracture aperture a of 10−3 m is used for the Chalk between Sandridge
and the Hatfield area. It is considered that there is a large uncertainty associated with this parameter, and
therefore, for the purposes of uncertainty estimation, the range is taken to be from an order of magnitude
lower to an order of magnitude higher.
There is little available local information for fracture spacing in the region of the source site and
1 km down-gradient. However, geophysical logging undertaken by TVW in the Hatfield area indicates
flowing fracture separations of approximately 1.00 m to 1.40 m. This is comparable to the flowing
fracture spacing observed from the single borehole dilution testing (Appendix D). These values fall at
the upper end of the range of the literature values for unweathered Chalk reviewed in Section 2.7.3.3.
Although, as noted in the review, these values relate to observed discontinuities, and may not relate to
the presence of hydraulically significant (flowing) fissures. The values relating to the spacing of flowing
horizons tend to have larger spacings. Therefore, for the purposes of uncertainty analysis, the range of
block sizes (2b) is taken to be from 0.50 to 1.50 m, with a mean of 1.00 m.
The darcy flux, q, was taken from the results of the borehole dilution testing at Nashe’s Farm,
Harefield House, and Comet Way (Section 3.6.1). This was converted to a velocity, v, using the mobile
porosity determined from a and b. This was then converted to a travel time ta for the path length. The
darcy fluxes ranged from 0.5 to 3.0 m day−1 , with average of 1.0 m day−1 (Nashe’s Farm), and 0.3 to
1.3 m day−1 , with average of 0.8 m day−1 (Harefield House). At Comet Way, darcy flux was almost an
order of magnitude higher: ranging from 4.5 to 14.5 m day−1 , with average of 8.0 m day−1 . Therefore,
for the branch Sandridge to Harefield House, and from Harefield House to Hatfield Quarry, q was taken
to be 0.3 m day−1 to 3.0 m day−1 , with an average of 1.0 m day−1 . For the branch Hatfield Quarry to
Comet Way q was taken to be 4.5 m day−1 to 14.5 m day−1 , with an average of 8.0 m day−1 .
The volumetric groundwater flux, Q, was determined from the recharge model used for the MOD-
FLOW modelling by Atkins (2004) and Cook (2010), as indicated by the MODPATH flowline. For
Sandridge to Hatfield, the recharge flux was 0.008 m day−1 , which over the site area of 7600 m2 at
SLC, results in a volumetric flux of 6.08 m3 day−1 . The volumetric flux was assumed to remain con-
stant along the flowline.
Table 6.1: Parameter combinations for ‘best-case’ (lowest peak bromate concentrations) and ‘worst-case’
(highest peak bromate concentrations) scenarios.
low concentrations.
Characteristic times are given in Table 6.2. The data for the breakthroughs from Harfield House and
Comet Way boreholes were not considered robust enough to be fit to the DP-1D model. Therefore, the
parameters tcb and σ for the branches from Harefield House and Comet Way to Arkley Hole and Lynch-
mill spring were taken to be the same as those from Water End to Arkley Hole and Lynchmill springs.
The values for ta were taken from the travel times determined by Cook (2010) for these connections.
Table 6.2: Parameters derived from fitting the DP-1D model (Barker, 2005) to tracer breakthrough curves
from Water End injection (Cook, 2010). Characteristic times are in hours.
Groundwater fluxes along the pathways were estimated from the tracer mass recovery data from
the tracer tests. Cook (2010) estimated the pro rata flow rate to each location based on the average
gauged flow to Water End swallow holes of 10756.8 m3 day−1 . The recovery percentage compared the
recovered mass to the amount available to recover at that time assuming literature derived values for
phage inactivation. The groundwater fluxes for the branches from Comet Way and Harefield House were
estimated by multiplying the recovery percentage for these connections by the groundwater flux from
the boreholes as estimated by the results from the single borehole dilution testing (Section 3.6.1.
For the branch Water End to Arkley Hole and Water End to Lynchmill Spring, the relative flux was
estimated by reference to the water balance by Cook (2010). Allogenic recharge accounted for 23.7 % of
flow and groundwater inflow for 14.4 %. Therefore, the flux in the karst system coming from Water End
was taken to be 165 % (23.7 % ÷ 14.4 %) of the flux along the flow line from SLC (i.e. 9.8 m3 day−1 )
and based on the tracer recovery data, this is apportioned as 7.2 m3 day−1 and 2.6 m3 day−1 to Arkley
Hole and Lynchmill respectively.
The value for α÷x was taken as 0.1 % for the branches from Water End, and 1.0 % for the branches
to the west of the main conduit network based from Harefield House and Comet Way.
served concentrations between 2000 and 2008. For Harefield House, the observed bromate concentra-
tions span the simulated bromate concentrations for Scenario A and Scenario C. For Hatfield Quarry and
Comet Way, Scenario C passes through the cluster of observed bromate concentrations, and simulated
concentrations for Scenario A are at the lower limits of observed bromate concentrations. For bromide,
the simulated concentrations for Scenario A pass through the cluster of observations at Harefield House,
Hatfield Quarry and Comet Way (although the simulated concentration ‘front’ appears to occur 5-10
years late at Hatfield Quarry), and simulated concentrations for Scenario B are well below observed
concentrations at all locations.
The 5000 µg l−1 constant concentration source term takes longer to reach peak concentrations,
which are also higher than for the other three source scenarios. The constant concentration source term
predicts bromate concentrations lower than observed between 2000 and 2008. This is in contrast to
Atkins (2005) and Cook (2010) who found that the same constant concentration source term gave good
agreement with the observed concentrations between 2000 and 2008, and is probably due to the increased
travel time from the source site to the monitoring locations introduced by the representation of vertical
migration through a low permeability ‘putty chalk’ layer at the source site in the modelling in this thesis.
The general form of the predictions is a bell-shaped curve, with an extended ‘tail’ at later times
as a result of the diffusion from matrix water back into the mobile water. The predicted peak bromate
concentrations occur first for Scenario C, then for Scenario A, and last for Scenario B which shows a
much lower and broader peak.
The importance of double-porosity diffusion in maintaining an elevated ’tail’ can be clearly seen
by comparing the simulated fissure and matrix concentrations at the end of each branch. For Scenario
A, matrix and fracture concentrations attain diffusive equilibrium at 2060, 2080 and 2085 for Harefield
House, Hatfield Quarry and Comet Way respectively. For Scenario C, matrix and fracture concentrations
attain diffusive equilibrium at 2045, 2065 and 2070 for Harefield House, Hatfield Quarry and Comet Way
respectively. After this time, bromate within the matrix provides a secondary source of contamination
which acts to maintain elevated concentrations in the fissures for a prolonged period of time.
At Harefield House, simulated node concentrations remain above 10 µg l−1 until 2225 for Scenarios
A and C and 2175 for Scenario B. At Hatfield Quarry, simulated node concentrations remain above
10 µg l−1 until 2325 for Scenarios A and C and 2250 for Scenario B. At Comet Way, simulated node
concentrations remain above 10 µg l−1 until 2310 for Scenarios A and C and 2230 for Scenario B.
The ‘worst-case’ and ‘best-case’ scenarios (as illustrated for Scenario C in Figure 6.13) affect the
timing, magnitude and sharpness of the peak bromate concentration. Compared to the‘typical-case’
Scenario, ‘worst-case’ scenarios have a narrower, higher concentration peak which occurs earlier, and
‘best-case’ scenarios have a broader, lower concentration peak which occurs later. The relative difference
between the magnitude and timing of predicted concentrations for best, worst and typical scenarios
increases as the distance from the source site increases. The differences in the magnitude and duration of
the bromate concentration peaks are due to the extent of diffusive exchange that occurs between fissure
and matrix. The most bromate mass diffuses into the immobile matrix porewater from the mobile fissure
6.7. Discussion and conclusions 256
water for the ‘best-case’ set of parameters. This attenuates the rise of bromate concentrations in the
fissures, but the back-diffusion from the matrix porewater to the fissures slows the falling limb of the
concentration peak.
For the spring concentrations, simulated bromate and bromide concentrations are significantly lower
(by around an order of magnitude) than observed concentrations. The concentrations at the spring nodes,
reflect the contributions from the four branches joining the spring node (Figure 6.8 and Figure 6.9),
along with the appropriate dilution at the node to represent water flux from flow lines that do not contain
bromate. The double-porosity branch contributes bromate concentrations in a broad peak; maximum
concentrations occur later (between 2050 and 2100) and are maintained for a longer duration, than peak
bromate concentrations provided by the karstic branches. The karstic branches from Harefield House
and Comet Way transport bromate at high concentrations within the fissures to the Arkley Hole and
Lynchmill Spring, while the karstic branch from the ‘karst junction’ to the springs, transports lower
concentrations, although the relative flux is much higher. Simulated matrix concentrations at the end of
the karst branches shows that double-porosity diffusion between matrix and fissure concentrations does
have a significant effect in attenuating bromate concentrations. However, substantially more diffusion of
bromate occurs along the double-porosity branch.
1.0x10-3
1.0x10-4 0.0
1950 2000 2050 2100 2150 2200 2250 2300 2350 1950 1970 1990 2010 2030 2050
End of Branch:
Harefield House SLC to Harefield House
3.0 3.0
Cnode Cf & Cm
Bromate concentation (mg l-1)
observed GW conc.
Sim. conc. Scenario A - Cnode
2.0 2.0
Sim. conc. Scenario A - Cm
Sim. conc. Scenario A - Cf
1.0 1.0
0.0
1950 2000 2050 2100 2150 1950 2000 2050 2100 2150
End of Branch:
Harefield House SLC to Harefield House
3.0 3.0
Cnode Cf & Cm
Bromate concentation (mg l-1)
observed GW conc.
Sim. conc. Scenario B - Cnode
2.0 2.0
Sim. conc. Scenario B - Cm
Sim. conc. Scenario B - Cf
1.0 1.0
0.0
1950 2000 2050 2100 2150 1950 2000 2050 2100 2150
End of Branch:
Harefield House SLC to Harefield House
3.0 3.0
Cnode Cf & Cm
Bromate concentation (mg l-1)
observed GW conc.
Sim. conc. Scenario C - Cnode
2.0 2.0
Sim. conc. Scenario C - Cm
Sim. conc. Scenario C - Cf
1.0 1.0
0.0
1950 2000 2050 2100 2150 1950 2000 2050 2100 2150
Figure 6.5: Simulated bromate concentrations at Harefield House using source terms for Scenario A, B
and C (Section 5.11), and a constant concentration source term of 5000 µg l−1 .
6.7. Discussion and conclusions 258
-3
1.0x10 0.4
-4
1.0x10 0.0
1950 2000 2050 2100 2150 2200 2250 2300 2350 1950 1970 1990 2010 2030 2050
End of Branch:
Hatfield Quarry Harefield House to Hatfield Quarry
2.0 2.0
Cnode Cf & Cm
Bromate concentation (mg l-1)
0.4 0.4
0.0
1950 2000 2050 2100 2150 1950 2000 2050 2100 2150
End of Branch:
Hatfield Quarry Harefield House to Hatfield Quarry
2.0 2.0
Cnode Cf & Cm
Bromate concentation (mg l-1)
0.4 0.4
0.0
1950 2000 2050 2100 2150 1950 2000 2050 2100 2150
End of Branch:
Hatfield Quarry Harefield House to Hatfield Quarry
2.0 2.0
Cnode Cf & Cm
Bromate concentation (mg l-1)
0.4 0.4
0.0
1950 2000 2050 2100 2150 1950 2000 2050 2100 2150
Figure 6.6: Simulated bromate concentrations at Hatfield Quarry using source terms for Scenario A, B
and C (Section 5.11), and a constant concentration source term of 5000 µg l−1 .
6.7. Discussion and conclusions 259
1.0x10-3 0.2
-4
1.0x10 0.0
1950 2000 2050 2100 2150 2200 2250 2300 2350 1950 1970 1990 2010 2030 2050
End of Branch:
Comet Way Hatfield Quarry to Comet Way
1.0 1.0
Cnode Cf & Cm
Bromate concentation (mg l-1)
0.2 0.2
0.0
1950 2000 2050 2100 2150 1950 2000 2050 2100 2150
End of Branch:
Comet Way Hatfield Quarry to Comet Way
1.0 1.0
Cnode Cf & Cm
Bromate concentation (mg l-1)
0.2 0.2
0.0
1950 2000 2050 2100 2150 1950 2000 2050 2100 2150
End of Branch:
Comet Way Hatfield Quarry to Comet Way
2.0 2.0
Cnode Cf & Cm
Bromate concentation (mg l-1)
0.4 0.4
0.0
1950 2000 2050 2100 2150 1950 2000 2050 2100 2150
Figure 6.7: Simulated bromate concentrations at Comet Way using source terms for Scenario A, B and
C (Section 5.11), and a constant concentration source term of 5000 µg l−1 .
6.7. Discussion and conclusions 260
observed GW conc.
Sim. conc. Scenario A
1.0x10-1 Sim. conc. Scenario B 0.03
Sim. conc. Scenario C
Sim. conc. Constant source
5 mg l-1 1965-2015
1.0x10-2 0.02
1.0x10-3 0.01
1.0x10-4 0.00
1950 2000 2050 2100 2150 2200 2250 1950 2000 2050
End of Branch: End of Branch:
Bromate concentation (mg l-1)
0.30
0.20
0.10
0.00
1950 2000 2050 2100 2150 1950 2000 2050 2100 2150
Bromate concentation (mg l-1)
0.50 End of Branch: Water End to Arkley Hole End of Branch: Water End to Arkley Hole
Cf Cm
0.40
0.30
0.20
0.10
0.00
1950 2000 2050 2100 2150 1950 2000 2050 2100 2150
Bromate concentation (mg l-1)
2.50 End of Branch: Harefield House to Arkley Hole End of Branch: Harefield House to Arkley Hole
Cf Cm
2.00
1.50
1.00
0.50
0.00
1950 2000 2050 2100 2150 1950 2000 2050 2100 2150
Bromate concentation (mg l-1)
2.50 End of Branch: Comet Way to Arkley Hole End of Branch: Comet Way to Arkley Hole
Cf Cm
2.00
1.50
1.00
0.50
0.00
1950 2000 2050 2100 2150 1950 2000 2050 2100 2150
Figure 6.8: Simulated bromate concentrations at Arkley Hole Spring node, and at the end of contributing
branches, using source terms for Scenario A, B and C (Section 5.11), and a constant concentration source
term of 5000 µg l−1 .
6.7. Discussion and conclusions 261
0.03
-2
1.0x10
0.02
-3
1.0x10
0.01
1.0x10-4 0.00
1950 2000 2050 2100 2150 2200 2250 1950 2000 2050
0.30
0.20
0.10
0.00
1950 2000 2050 2100 2150 1950 2000 2050 2100 2150
Bromate concentation (mg l-1)
0.50 End of Branch: Water End to Lynchmill End of Branch: Water End to Lynchmill
Cf Cm
0.40
0.30
0.20
0.10
0.00
1950 2000 2050 2100 2150 1950 2000 2050 2100 2150
Bromate concentation (mg l-1)
2.50 End of Branch: Harefield House to Lynchmill End of Branch: Harefield House to to Lynchmill
Cf Cm
2.00
1.50
1.00
0.50
0.00
1950 2000 2050 2100 2150 1950 2000 2050 2100 2150
Bromate concentation (mg l-1)
2.50 End of Branch: Comet Way to Lynchmill End of Branch: Comet Way to Lynchmill
Cf Cm
2.00
1.50
1.00
0.50
0.00
1950 2000 2050 2100 2150 1950 2000 2050 2100 2150
Figure 6.9: Simulated bromate concentrations at Lynchmill Spring node, and at the end of contributing
branches, using source terms for Scenario A, B and C (Section 5.11), and a constant concentration source
term of 5000 µg l−1 .
6.7. Discussion and conclusions 262
6.0
1.0x10-1
4.0
-2
1.0x10
2.0
-3
1.0x10 0.0
1950 2000 2050 2100 2150 2200 2250 2300 2350 1950 1970 1990 2010 2030 2050
End of Branch:
Harefield House SLC to Harefield House
10.0 10.0
Cnode Cf & Cm
Bromide concentation (mg l-1)
2.0 2.0
0.0
1950 2000 2050 2100 2150 1950 2000 2050 2100 2150
End of Branch:
Harefield House SLC to Harefield House
10.0 10.0
Cnode Cf & Cm
Bromide concentation (mg l-1)
2.0 2.0
0.0
1950 2000 2050 2100 2150 1950 2000 2050 2100 2150
Figure 6.10: Simulated bromide concentrations at Harefield House using source terms for Scenario A
and B.
6.7. Discussion and conclusions 263
3.0
-1
1.0x10
2.0
1.0x10-2
1.0
-3
1.0x10 0.0
1950 2000 2050 2100 2150 2200 2250 2300 2350 1950 1970 1990 2010 2030 2050
End of Branch:
Hatfield Quarry Harefield House to Hatfield Quarry
5.0 5.0
Cnode Cf & Cm
Bromide concentation (mg l-1)
1.0 1.0
0.0
1950 2000 2050 2100 2150 1950 2000 2050 2100 2150
1.0 1.0
0.0
1950 2000 2050 2100 2150 1950 2000 2050 2100 2150
Figure 6.11: Simulated bromide concentrations at Hatfield Quarry using source terms for Scenario A
and B.
6.7. Discussion and conclusions 264
3.0
-1
1.0x10
2.0
1.0x10-2
1.0
-3
1.0x10 0.0
1950 2000 2050 2100 2150 2200 2250 2300 2350 1950 1970 1990 2010 2030 2050
End of Branch:
Comet Way Hatfield Quarry to Comet Way
5.0 5.0
Cnode Cf & Cm
Bromide concentation (mg l-1)
observed GW conc.
4.0 Sim. conc. Scenario A - Cnode 4.0
Sim. conc. Scenario A - Cm
3.0 3.0
Sim. conc. Scenario A - Cf
2.0 2.0
1.0 1.0
0.0
1950 2000 2050 2100 2150 1950 2000 2050 2100 2150
End of Branch:
Comet Way Hatfield Quarry to Comet Way
5.0 5.0
Cnode Cf & Cm
Bromide concentation (mg l-1)
1.0 1.0
0.0
1950 2000 2050 2100 2150 1950 2000 2050 2100 2150
Figure 6.12: Simulated bromide concentrations at Comet Way using source terms for Scenario A and B.
6.7. Discussion and conclusions 265
observed GW conc.
Sim. Cm. Scenario C - BEST
Sim. Cnode or Cf. Scenario C - BEST
Sim. Cm. Scenario C - WORST
Sim. Cnode or Cf. Scenario C - WORST
Sim. Cm. Scenario C - TYPICAL
Sim. Cnode or Cf. Scenario C - TYPICAL
Harefield House
3.0 3.0
Cnode End of Branch:
SLC to Harefield House
Bromate concentation (mg l-1)
Cf & Cm
2.0 2.0
1.0 1.0
0.0
1950 2000 2050 2100 2150 2200 2250 1950 2000 2050 2100 2150 2200 2250
2.0 2.0
1.0 1.0
0.0 0.0
1950 2000 2050 2100 2150 2200 2250 1950 2000 2050 2100 2150 2200 2250
2.0 2.0
1.0 1.0
0.0 0.0
1950 2000 2050 2100 2150 2200 2250 1950 2000 2050 2100 2150 2200 2250
Figure 6.13: Simulated bromate concentrations for Scenario C at Harefield House, Hatfield Quarry, and
Comet Way using ‘best-case’, ‘typical-case’ and ‘worst-case’ parameters.
6.7. Discussion and conclusions 266
in a number of ways: the point of time in the contaminant evolution which is represented by the obser-
vations is not evident. A longer period of monitoring data would be necessary to identify trends which
could allow the curve to be calibrated with more certainty. Furthermore, observations of a particular
fissure concentration can represent one of two points on the simulated curve (Figure 6.14). Concurrent
observations of matrix porewater concentrations are required to identify which point in the evolution
such a fissure concentration represents.
Figure 6.14: Concurrent matrix and fissure concentrations are required to determine at which point along
the concentration-time graph a particular fissure concentration represents.
The network model has been less successful in representing concentrations at locations to the east
of the Palaeogene escarpment, where the karst flow system is believed to have a dominant effect on the
migration of bromate contamination across the catchment. This could be due to the uncertainty in the
relative fluxes of the karst system. More extensive tracer testing and flow gauging could help to reduce
this uncertainty.
The representation of the karst branches within the network model is a very simplistic approxima-
tion to the likely reality. The karst branches represented are only those that were identified by tracer
testing, but in reality there are likely to be additional connections. Also, very limited interaction between
karst branches and double-porosity branches is simulated by this model: the flow pathway between the
Hatfield Area and Arkley Hole and Lynchmill Spring is represented by either a single karst branch or
a single double-porosity branch. Further work could experiment with more extensive branching and
joining between karst branches and double-porosity branches.
The Network model represents a one-dimensional steady-state flow regime between nodes. There-
fore, unlike numerical models such as MODFLOW and MT3D, the network model model output does
not capture seasonal variations in concentrations, nor incorporate transient flow conditions such as those
6.7. Discussion and conclusions 267
Chapter 7
Conclusions
This chapter considers the contributions of the analysis and interpretation of data and the modelling work
presented in Chapters 3, 4, 5 and 6 in meeting the aims and objectives of the research outlined in Chap-
ter 1. For convenience, the research objectives are repeated and discussed in sections 7.1.1, 7.1.2 to 7.1.3.
The research presented in this thesis has fulfilled this overall aim by developing a refined conceptual
model of bromate transport within the catchment and interpreting the spatial distribution of bromate in
light of this conceptual understanding, by conceptualising and quantifying a range of source history
scenarios for bromate input to the aquifer, and by using these advancements in understanding as a basis
for developing a network model that demonstrates the influences of double-porosity diffusion on the
long-term evolution of bromate at a catchment-scale.
• To develop a conceptual model for groundwater flow and contaminant transport in the Hertford-
shire Chalk aquifer system by review of existing data and interpretation of additional tracer testing
and geophysical testing;
• To use the available information and monitoring data to describe the spatial distribution and
temporal evolution of bromate across the catchment, and to interpret this in association with the
conceptual model of the flow and transport system.
The detailed analysis of bromate and bromide monitoring data presented in Chapter 4 has revealed
that bromate concentrations are affected by influences including recharge (soil moisture deficit, rainfall),
7.1. Fulfillment of research aims and objectives 269
water level, and abstractions. These relationships, integrated with the observations of the geology, hy-
drogeology and hydrology of the area affected by the bromate contamination, have been used to refine
the conceptual model of groundwater flow and transport of bromate within the catchment Chapter 3.
The conceptualisation supports double-porosity dominated transport of bromate within the Vale of St.
Albans area, which maintains a highly attenuated, relatively stable contaminant distribution west of
Hatfield. To the east of Hatfield, a significant karst network related to the position of the Palaeogene
overlap of the Chalk influences bromate trasport, dispersing bromate over large distances toward the
northern and middle Lea Valley. The revised conceptual understanding provides the basis for modelling
approaches applied to predict long-term, large-scale transport of bromate within the Hertfordshire Chalk
(section 7.1.3), and has allowed a new interpretation of the spatial distribution and evolution of bromate
and bromide within the catchment to be developed between 2000 and 2008.
However, the interpretation of the spatial and temporal evolution of bromate and bromide within the
catchment is hampered by a number of inadequacies in the available monitoring data: trends are difficult
to discern because monitoring data is available for a relatively short period of time, monitoring frequency
varies considerably between locations and varies over time at individual locations, and there are strong
seasonal influences. Data are generally for (non depth-specific) pumped groundwater samples so that
vertical distribution of bromate contamination cannot be investigated, nor can the matrix porewater con-
centrations be diagnosed. This is a severe limitation of the monitoring and investigation programme,
effectively preventing proper consideration of diffusive retardation of bromate which the thesis shows
may prolong the occurrence of bromate contamination by 200 years.
• To describe and quantify the distribution of bromate at the source site through collation and de-
scription of site investigation and monitoring data;
• To develop alternative conceptual scenarios for bromate release to groundwater and quantify these
as ‘source terms’;
• To use the available monitoring data to constrain the potential source terms.
In Chapter 5, the available site investigation data was assessed and interpreted to estimate the quan-
tity of bromate and bromide present on site in the unsaturated and saturated zone soils and groundwater
prior to the redevelopment of the site in the mid 1980s and subsequent to the discovery of bromate con-
tamination in early 2000s. These estimates have been used to constrain three source term scenarios for
bromate input to the aquifer beneath the site. There are many uncertainties associated with an incom-
plete knowledge of the history of the site, and the three source scenarios attempt to capture the range
of possible bromide and bromate source histories. This rigorous analysis of the source zone provides
a significant improvement in the characterisation of the bromate ‘source term’ compared to previous
representations, particularly the constant concentrations source terms used by Buckle (2003) and Atkins
7.1. Fulfillment of research aims and objectives 270
(2005). Nevertheless, it is recognised that due to the relative scarcity of data, these scenarios still repre-
sent crude estimates of the reality.
The one-dimensional double-porosity transport code, DP1D (Barker, 2005), has been used to simu-
late concentrations in groundwater down-gradient of the source site. Simulated concentrations using two
of the source term scenarios show relatively good agreement with observed groundwater concentrations
at locations 150 m, 500 m and 1000 m down-gradient of the source site. Either a ‘catastrophic release’
of bromide/bromate to the unsaturated zone followed by leaching to groundwater beneath the site, or a
‘direct release’ of bromate to the saturated zone, sometime between 1960 and 1970, result in the closest
fit to observed data. Porewater concentrations are not available for locations down-gradient of the source
site. The lack of porewater concentrations and the relatively short period of time for which groundwater
monitoring data (fissure concentrations) are available, combined with the large seasonal variations in
concentrations, means that the trends are difficult to discern, and robust conclusions cannot be made as
to whether or not the simulations are representative of the actual release scenarios.
• To use this model to produce predictions for the likely bromate concentrations at key output loca-
tions over the long-term.
A novel analytical network model to represent the Hertfordshire Chalk catchment has been devel-
oped in Chapter 6, using code written by Prof. John Barker. The network model simulates Fickian
double-porosity diffusive exchange along interconnecting flow-lines, while allowing karstic branches to
be incorporated into the network. The model was parameterised by a combination of values found within
the literature, and the results of the single borehole dilution testing and catchment-scale natural gradient
tracer testing.
The network model, using the range of source terms developed in Chapter 5, has been successful in
simulating bromate and bromide concentrations of the order of magnitude of those observed at locations
within the Vale of St. Albans, west of the main karst system. The network model highlights the long-term
effects of double-porosity nature of the chalk on the catchment-scale evolution of bromate. The double-
porosity diffusion significantly extends the duration of elevated levels of contamination by providing a
secondary source of bromate: bromate concentrations within the Vale of St. Albans are predicted to
remain above regulatory limits for around 200 years.
The network model has been less successful in representing concentrations at locations to the east
of the Palaeogene escarpment, where the karst flow system is believed to have a dominant effect on the
migration of bromate contamination across the catchment. This is largely due to the uncertainty in the
relative fluxes of the karst system. The representation of the karst branches within the network model is
a very simplistic approximation to the likely reality. The karst branches represented are only those that
7.2. Recommendations for further work 271
were identified by tracer testing, and very limited interaction between karst branches and double-porosity
branches is simulated by this model. Further work could experiment with more extensive branching of
the karst network and closer interaction between karst branches and double-porosity branches.
The Network model represents a one-dimensional steady-state flow regime between nodes. There-
fore, unlike numerical models such as MODFLOW and MT3D, the network model model output does
not capture seasonal variations in concentrations, nor incorporate transient flow conditions such as those
introduced by changes in abstraction regimes. However, the analytical network model has a number of
advantages over the numerical models. In particular, the analytical model is computationally fast over
long time-scales, taking seconds to minutes to run compared to days to weeks for the catchment MODL-
FOW and MT3D-MS models (Atkins, 2005; Cook, 2010). Also, for the analytical network model, the
number of parameters which are varied during the calibration process is small compared to a distributed
parameter numerical model. No additional calibration was undertaken on the results within this thesis
due to insufficient monitoring data. Without measured bromate concentrations in porewater and longer
time-series of bromate concentrations in fissure water becoming available, no effective calibration will
be possible.
Bibliography
Allen, D. J., Brewerton, L. J., Coleby, L. M., Gibbs, B. R., Lewis, M. A., MacDonald, A. M., Wagstaff,
S. J. and Williams, A. T. (1997), The physical properties of major aquifers in England and Wales.
British Geological Survey Technical Report WD/97/34, Technical Report, British Geological Survey.
Atkins (2002), St. Leonard’s Court, Sandridge, St. Albans. Environmental Site Investigation and Quan-
titative Pollutant Linkage Assessment., Technical Report, Atkins Ltd.
Atkins (2004), Bromate Contamination in the Lee Valley. Phase 1: Data Collation and Conceptualisa-
tion., Technical Report, Atkins Limited.
Atkins (2005), Bromate Contamination in the Lee Valley Phase 2: Modelling Report., Technical Report,
Atkins Limited.
Atkins (2006), Bromate Monitoring Data Review, Technical Report, Atkins Limited.
Atkinson, T. C. and Smith, D. I. (1974), Rapid groundwater flow in fissures in the chalk: an example
from south Hampshire, Quarterly Journal of Engineering Geology and Hydrogeology 7(2), 197–205.
Atkinson, T. and Smart, P. (1981), Artificial Tracers in Hydrology, in A Survey of British Hydrology,
Royal Society, London, pp.173–190.
Banks, D., Davies, C. and Davies, W. (1995), The Chalk as a karstic aquifer: evidence from a tracer test
at Stanford Dingley, Berkshire, UK, Quarterly Journal of Engineering Geology and Hydrogeology
28(Supplement1 ), S31 − −38.
Barker, J. A. (1982), Laplace transform solutions for solute transport in fissured aquifers, Advances in
Water Resources 5(2), 98–104.
Barker, J. A. (1985b), Modelling the effects of matrix diffusion on transport in densely fissured media., in
Hydrogeology in the Service of Man. Memoirs of the 18th Congress of the International Association fo
Hydrogeologists., Vol. XVIII, International Association of Hydrogeologists, Cambridge, pp.250–269.
Barker, J. A. (1993), Modelling of flow and transport in the chalk, in R. A. Downing, M. Price and G. P.
Jones (eds.), The Hydrogeology of the Chalk of North-West Europe, Clarendon Press, Oxford, pp.35–38.
Barker, J. A. (2001), Mathemetical basis for the MAP model., Technical Report.
Barker, J. A. (2005), DP1D-Link Description of the code and issues. Version 6., Technical Report.
Barker, J. A. and Foster, S. S. D. (1981), A Diffusion Exchange Model for Solute Movement in Fissured
Porous Rock, Quarterly Journal of Engineering Geology 14(1), 17–24.
Barker, J. A., Wright, T. and Fretwell, B. (2000), A pulsed-velocity method of solute transport modelling.,
in A. Dassargues (ed.), Tracers and Modelling in Hydrogeology, Proceedings of the TraM’2000 Con-
ference, Vol. International Association of Hydrological Sciences Publication 262, IAHS Press, Liege,
Belgium, p.297302.
Barten, W. (1996), Linear response concept combining advection and limited rock matrix diffusion in a
fracture network transport model, Water Resources Research 32(11), 3285–3296.
Bibby, R. (1981), Mass treansport studies in dual porosity media., Water Resources Research 17, 1075–
1081.
Bloomfield, J. P., Brewerton, L. J. and Allen, D. J. (1995), Regional trends in matrix porosity and dry
density of the chalk of England, Quarterly Journal of Engineering Geology 28, S131–S142.
Bloomfield, J. P., Farrant, A. R. and Cunningham, J. (2004), Structural Interpretation of the Chalk of the
Hertford District based on Slope Aspect Mapping (Draft). BGS Report Ref. CR/04/XXC, Technical
Report Ref. CR/04/XXC, British Geological Survey.
Board, W. R. (1972), The Hydrogeology of the London Basin, Technical Report, Water Resources Board.
Bristow, R., Mortimore, R. and Wood, C. (1997), Lithostratigraphy for mapping the Chalk of southern
England, Proceedings of the Geologists Association 108, 293–315.
Buckle, D. (2002), Bromate Groundwater Flow Study. Phase 1 (Conceptual Understanding) Report. Under-
taken on behalf of the Environment Agency and Three Valleys Water., Technical Report, Vivendi Water
Partnership.
Buckle, D. (2003), Bromate Groundwater Flow Study. Phase 2 Interim (Modelling) Report. Undertaken on
behalf of the Environment Agency and Three Valleys Water., Technical Report, Vivendi Water Partner-
ship.
BIBLIOGRAPHY 274
Burgess, W., Barker, J., Watson, S. and Fretwell, B. (2005), Contaminant retardation by double porosity
diffusive exchange in the Chalk: implications for monitoring., in Bringing Groundwater Quality Re-
search to the Watershed scale. Proceedings of the 4th International Groundwater Quality Conference..,
IAH, Waterloo, Canada.
Butler, R., Godley, A., Lytton, L. and Cartmell, E. (2005), Bromate environmental contamination: Review
of impact and possible treatment, Critical Reviews in Environmental Science and Technology 35(3), 193–
217.
Chatfield, C. (2004.), The Analysis of Time Series: An Introduction., sixth edition edition, Chapman and
Hall/CRC.
Chemfix (1985a), Evaluation of the results from the borehole situated 120m down dip from the Sandridge
site., Technical Report.
Chemfix (1985b), Field report for drilling of borehole C1 at House Lane, Sandridge on 21/22 January 1985,
Technical Report.
Chemfix (1985c), Report of the second phase of the field investigation at the Sandridge site, Technical
Report.
Chemfix (1985d), Report on the further modelling studies for the Sandridge site., Technical Report.
CL:AIRE (2002), Site characterisation in support of monitored natural attenuation of fuel hydrocarbons
and MTBE in a chalk aquifer in southern England., in Case study bulletin CSB1., Contaminated Land:
Applications in Real Environments (CL:AIRE).
Cook, S. J. (2010), The hydrogeology of bromate contamination in the Hertfordshire Chalk: Incorporating
Karst in Predictive Models., EngD, University College London.
Domenico, P. and Schwartz, W. (1998), Physical and Chemical Hydrogeology., second edition edition,
Wiley.
EA (2005), Part IIA of the Rnvironmental Protection Act 1990. Decision Document for St Leonard’s Court.,
Technical Report, Environment Agency.
Edmunds, W. M., Cook, J. M., Kinniburgh, D. G., Miles, D. L. and Trafford, J. M. (1989), Trace-element
occurrence in British groundwaters., Technical Report BGS Research Report SD/89/3, British Geologi-
cal Survey.
Edwards, A. J., Hobbs, S. L. and Smart, P. L. (1991), Effects of quarry dewatering on a karstified limestone
aquifer: A case study from the Mendip Hills, England., in Third Conference on Hydrogeology, Ecology,
Monitoring, and Management of Ground Water in Karst Terranes., National Water Well Association,
pp.77–92.
Entec (2000), River Mimram and Upper Lee Water Resources Sustainability Study Groundwater Model.
Phase 1 Report: Data Collation and Formulation of Conceptual Model. Undertaken on behalf of the
Environment Agency., Technical Report, Entec.
Entec (2002), River Mimram and Upper Lee Water Resources Sustainability Study Groundwater Model.
Phase 2 Model Development and Refinement. Undertaken on behalf of the Environment Agency., Tech-
nical Report, Entec.
Fitzpatrick, C. M. (2007), Hatfield Pumping Test. Interim Report for TVW and TWUL for the period July
2005 to December 2006., Technical Report, University College London.
Ford, D. C. and Williams, P. (2007), Karst hydrogeology and geomorphology, John Wiley, Chichester.
Foster, S. S. D. (1975), The Chalk groundwater tritium anomaly – A possible explanation, Journal of
Hydrology 25(1-2), 159–165.
Foster, S. S. D. and Milton, V. A. (1974), The permeability and storage of an unconfined chalk aquifer,
Hydrological Sciences Bulletin 19, 485–500.
Gale, A., Hancock, J., Bristow, R., Mortimore, R. and Wood, C. (1999), ’Lithostratigraphy for mapping
the Chalk of southern England’ by Bristow et al. (1997): discussion, Proceedings of the Geologists
Association 110, 65–72.
GoldsimTechnologyGroup (2007), GoldSim Contaminant Transport Module. User Guide, Technical Re-
port, GoldsimTechnologyGroup.
Goudie, A. (1990), The Landforms of England and Wales, Basil Blackwell, Oxford.
Grisak, G. E. and Pickens, J. F. (1980), Solute Transport through Fractured Media .1. The Effect of Matrix
Diffusion, Water Resources Research 16(4), 719–730.
Grisak, G. E. and Pickens, J. F. (1981), An Analytical Solution for Solute Transport through Fractured
Media with Matrix Diffusion, Journal of Hydrology 52(1-2), 47–57.
Hancock, J. M. (1975), The petrology of the Chalk, Proceedings of the Geologists’ Association 86(4), 499–
535.
BIBLIOGRAPHY 276
Harbaugh, A. W., Banta, E. R., Hill, M. and McDonald, M. (2000), MODFLOW-2000, The U.S. Geological
Survey modular ground-water modelUser Guide to modularization concepts and the ground-water flow
process., Technical Report, U.S. Geological Survey,.
Harold, C. H. H. (1937), Thirty-second Annual Report on the results of the chemcial and bacteriological
examination of the London waters for the twelve months ended 31st December 1937., Technical Report,
Metropolitan Water Board.
Headworth, H. (1972), The analysis of natural groundwater fluctuations in the chalk of Hampshire., Journal
of the Institute of Water Engineers 26, 107–124.
Headworth, H. G., Keating, T. and Packman, M. J. (1982), Evidence for a shallow highly-permeable zone
in the Chalk of Hampshire, U.K, Journal of Hydrology 55(1-4), 93–112.
Hijnen, W., Voogt, R., Veenendaal, H. and van der Jagt, H. (1995), Bromate reduction by denitrifying
bacteria., Applied Environmental Microbiology 61, 239.
Hill, D. (1984), Diffusion coefficients of nitrate, chloride, sulphate and water in cracked and uncracked
Chalk., Journal of Soil Science 35, 27–33.
Hydrotechnica (1988), Northern New River Wellfield Pump Testing. Final Report., Technical Report,
Hydrotechnica.
Ineson, J. (1962), A hydrogeological study of the permeability of Chalk, Journal of the Institution of Water
Engineers 16, 449–463.
Kachi, S. (1987), Tracer studies in the Chalk aquifer near Cambridge., PhD thesis, University of East
Anglia.
Klinck, B. A., Hopson, P. M., Lewis, M. A., MacDonald, D. M., Inglethorpe, S. D. J., Entwisle, D. C., Har-
rington, J. F. and Williams, L. (1998), The hydrogeological behaviour of the clay-with-flints in southern
England. British Geological Survey Technical Report, WE/97/5., Technical Report, British Geological
Survey.
Komex (2000), Site Investigation at St Leonards Court, Sandridge, St Albans., Technical Report 50598,
Komex.
Kruithof, J. and Meijers, R. (1995), Bromate formation by ozonation ans advanced oxidation and potential
options in drinking water treatment., Water Supply 13, 93.
BIBLIOGRAPHY 277
Lewin, K., Bradshaw, K., Blakey, N., Turrell, J., Hennings, S. and Fleming, R. (1994), Leaching Tests for
Assessment of Contaminated Land: Interim NRA Guidance. National Rivers Authority RD Note 301 .,
Technical Report, National Rivers Authority.
MacDonald, A., Brewerton, L. and Allen, D. (1998), Evidence for rapid groundwater flow and karst
type behaviour in the Chalk of Southern England., in N. Robins (ed.), Groundwater Pollution, Aquifer
Recharge and Vulnerability, Geological Society of London Special Publication No. 130, pp.95–106.
MacDonald, A. M. and Allen, D. J. (2001), Aquifer properties of the Chalk of England, Quarterly Journal
of Engineering Geology and Hydrogeology 34(4), 371–384.
Mathias, S. A., Butler, A. P., Jackson, B. M. and Wheater, H. S. (2006), Transient simulations of flow and
transport in the Chalk unsaturated zone, Journal of Hydrology 330(1-2), 10–28.
Mathias, S. A., Butler, A. P., McIntyre, N. and Wheater, H. S. (2005), The significance of flow in the matrix
of the Chalk unsaturated zone, Journal of Hydrology 310(1-4), 62–77.
Maurice, L. D., Atkinson, T. C., Barker, J. A., Bloomfield, J. P., Farrant, A. R. and Williams, A. T. (2006),
Karstic behaviour of groundwater in the English Chalk, Journal of Hydrology 330(1-2), 63–70.
Mortimore, R. N. (1993), Chalk water and engineering geology., in R. A. Downing, M. Price and G. P.
Jones (eds.), The Hydrogeology of the Chalk of North-West Europe, Clarendon Press, Oxford, pp.67–92.
Mortimore, R., Pomerol, B. and Foord, R. J. (1990), Engineering stratigraphy and paeleogeography for
the chalk of the Anglo-Paris Basin, in Proceedings of the International Chalk Symposium 1989, Thomas
Telford.
Oakes, D. (1977), The movement of water and solutes through the unsaturated zone of the Chalk of the
United Kingdom., in 3rd International Hydrology Symposium, Fort Collins, Colorado State University,
pp.447–459.
Ogata, A. and Banks, R. B. (1961), A solution of the differential equation of longitudinal dispersion in
porous media., United States Geological Survey Professional Paper 411-A.
Pollock, D. W. (1989), Documentation of computer programs to compute and display pathlines using
resultsd form the U.S. Geological survey modular three-dimensional finite-difference ground-water
model., Technical Report, U.S. Geological Survey.
Price, M. (1987), Fluid flow in the Chalk of England, in J. C. Goff and B. P. J. Williams (eds.), Fluid Flow
in Sedimentary Basins and Aquifers. Geological Society of London Special Publication 34, pp.141–156.
Price, M., Atkinson, T. C., Barker, J. A., Wheeler, D. and Monkhouse, R. A. (1992), A Tracer Study of the
Danger Posed to A Chalk Aquifer by Contaminated Highway Run-Off, Proceedings of the Institution of
Civil Engineers-Water Maritime and Energy 96(1), 9–18.
BIBLIOGRAPHY 278
Price, M., Atkinson, T. C., Wheerler, D., Barker, J. A. and Monkhouse, R. A. (1989), The Movement
of Groundwater in the Chalk Aquifer near Bricket Wood, Hertfordshire, and its Possible Pollution
by Drainage from the M25 Motorway. British Geological Survey Report WD/89/3., Technical Report,
British Geological Survey.
Price, M., Bird, M. J. and Foster, S. S. D. (1976), Chalk pore-size measurements and their significance,
Water Services 80(968), 596–600.
Price, M., Downing, R. A. and Edmunds, W. M. (1993), The Chalk as an aquifer, in R. A. Downing,
M. Price and G. P. Jones (eds.), The Hydrogeology of the Chalk of North-West Europe, pp.35–58.
Price, M., Low, R. G. and McCann, C. (2000), Mechanisms of water storage and flow in the unsaturated
zone of the Chalk aquifer, Journal of Hydrology 233(1-4), 54–71.
Price, M., Morris, B. and Robertson, A. (1982), A study of intergranular and fissure permeability in Chalk
and Permian aquifers, using double-packer injection testing, Journal of Hydrology 54(4), 401–423.
Rawson, P. F., Allen, P. M. and Gale, A. S. (2001), The Chalk Group - a revised lithostratigraphy., Geosci-
entist 11, 21.
Roberts, M. (2001), The Sandridge/Hatfield Bromate Pollution Investigation, Technical Report, Environ-
ment Agency.
Rouleau, A. and Gale, J. E. (1985), Statistical characterisation of the fracture system in the Stripa Granite,
Sweden, Int. J. Rock Mech. Min. Sci. Geomech. Abstr. 22, 353–367.
Shand, P., Tyler-Whittle, R., Besien, T., Peach, D. W., Lawrence, A. R. and Lewis, H. O. (2003), Baseline
Report Series: 6. The Chalk of the Colne and Lee river catchments, Technical Report British Geological
Survey Commissioned Report CR/03/069N. Environment Agency National Groundwater Contaminated
Land Centre Technical Report NC/99/74/6, British Geological Survey and Environment Agency.
Silgram, M., Williams, A., Waring, R., Neumann, I., Hughes, A., Mansour, M. and Besien, T. (2005),
Effectiveness of the nitrate sensitive areas scheme in reducing groundwater concentrations in England,
Quarterly Journal of Engineering Geology and Hydrogeology 38, 117–127.
Smith, D., Wearn, P., Richards, H. and Rowe, P. (1970), Water movement in the unsaturated zone of high
and low permeabilitystrata using natural tritium., in Isotope Hydrology, Atomic Energy Agency, Vienna,
pp.73–87.
Snow, D. T. (1968), Rock fracture spacing, opening and porosities, Journal of the Soil and Foundations
Division, Proceedings of the American Society of Civil Engineers 94, 19.
STATS (1983a), Interim report on site investigation, ref 83/3105., Technical Report, St Albans Testing
Services.
BIBLIOGRAPHY 279
STATS (1983b), Second report on site investigation, ref 83/3105A, Technical Report, St Albans Testing
Services.
STATS (1983c), Third report on site investigation, ref 83/3105C, Technical Report, St Albans Testing
Services.
STATS (1984), Report on further site investigations. ref AM/3554, Technical Report, St Albans Testing
Services.
Thomas, J. (2000), Note on the 1984 and 1985 modelling studies., Technical Report, Environment Agency.
Walsh, P. and Ockenden, A. (1982), Hydrogeological observations at the Water End swallow hole complex,
North Mimms, Hertfordshire., Transactions of the British Cave Research Association. 9, 184–194.
Ward, R. S. (1989), Artificial tracer and natural 222Rn studies of the East Anglian Chalk aquifer., PhD
thesis, University of East Anglia, Norwich.
Ward, R. S., Williams, A. T., Barker, J. A., Brewerton, L. J. and Gale, I. N. (1998), Groundwater tracer
tests: a review and guidelines for their use in British aquifers. RD Technical Report W160, Technical
Report, Environment Agency. NOT IN FILE.
Watson, S. J. (2004), Solute transport and hydrodynamic characteristics in the Chalk aquifer at Tilmanstone,
Kent, PhD, University College London.
White, W. B. (2003), Conceptual models for karstic aquifers., Speleogenesis and Evolution of Karst
Aquifers 1(1), 1–6.
Whittaker, H. (1921), Water Supply of Buckinghamshire and Hertfordshire. Geological Survey Memoirs.,
Technical Report, British Geological Survey.
Williams, A., Barker, J. A., Silgram, M., Mansour, M., Neumann, I. and Hughes, A. (2003), The use of
flow modelling to assess the impact of agricultural control measures on abstracted groundwater quality.,
in MODFLOW and More. Understanding through Modelling., pp.446–450.
Williams, A., Bloomfield, J., Griffiths, K. and Butler, A. (2006), Characterising the vertical variations in
hydraulic conductivity within the Chalk aquifer, Journal of Hydrology 330(1-2), 53–62.
Williams, R. (1987), Frost weathered mantles on the Chalk., in J. Boardman (ed.), Periglacial processes
and landforms in Britain and Ireland, Cambridge University Press, Cambridge, pp.127–133.
Woods, M. A. (2003), The Stratigraphy of the Chalk Group of the Hertford District: a preliminary study
for Veolia Water Partnership. BGS Report Ref.CR/03/216, Technical Report Ref.CR/03/216, British
Geological Survey.
Woods, M. A. and Aldiss, D. T. (2004), The stratigraphy of the Chalk Group of the Berkshire Downs,
Proceedings of the Geologists Association 115, 249–265.
Worthington, S. . (2003), A comprehensive strategy for understanding flow in carbonate aquifers., Speleo-
genesis and Evolution of Karst Aquifers 1(1), 1–8.
Worthington, S., Ford, D. C., Beddows, P. A. and Dreybrodt, W. (2000), Porosity and permeability en-
hancement in unconfined aquifers as a result of solution., in A. Klimchouk, D. C. Ford and A. N. Palmer
(eds.), Speleogenesis: Evolution of karst aquifers., National Speleological Society, Huntsville, Alabama,
pp.463–472.
Younger, P. L. and Elliot, T. (1995), Chalk fracture system characteristics: implications for flow and solute
transport, Quarterly Journal of Engineering Geology and Hydrogeology 28(Supplement1 ), S39 − −50.
Zheng, C. (1990), MT3D: A modular three-dimensional transport model for simulation of advection, dis-
persion, and chemical reactions of contaminants in groundwater systems. Report to the U.S. Environ-
mental Porectection Agency., Technical Report.
Zheng, C. and Wang, P. P. (1999), MT3DMS: A modular three-dimensional multispecies transport model
for simulation of advection, dispersion, and chemical reactions of contaminants in groundwater systems;
documentation and users guide. Contract Report SERDP-99-1., Technical Report, U.S. Army Engineer
Research and Development Center.
281
Appendix A
This chronology of key events in the history of bromate contamination in Hertfordshire has been com-
piled based on information obtained from the Jon Newton at the Environment Agency, Rob Sage at
VWTV and Philip Bishop at TWUL.
1955 Planning permission granted for use of the site for the manufacture of specified chemicals.
1955–1980 The site occupied by Steetly Chemical Works and operated for the production of bromine-
based chemicals, including bromate.
1983 Crest Nicholson Residential Ltd (Crest) purchase the factory site and adjoining land.
1984 Crest are granted planning permission for 30 houses and commence demolition and clearance of
buildings and hardstanding from the site.
1983–1985 Intrusive investigations on behalf of Crest. The site is found to be contaminated with or-
ganic bromide compounds. The potential for bromide pollution of groundwater is recognised and
assessed by consultants to Crest.
1985–1986 Crest granted increased planning permission for 66 houses on the site. Approval requires
the removal of contaminated soil to minimum of 1 m depth across the entire site. Excavations
completed by August 1986.
1986–1987 Construction of the residential development St Leonard’s Court (SLC) begins in November
1986 and is completed by October 1987.
1998–1999 VWTV detect bromate at Essendon PWS at concentrations of 10.8 µg l−1 in December
1999. The EA are informed.
2000 VWTV detect bromate concentrations of ∼100 µg l−1 , well in excess of the future drinking water
standard of 10 µg l−1 , at Hatfield PWS. The EA are informed and abstractions for public supply
are ceased. A sampling programme is undertaken that identifies St Leonard’s Court as the source
of the contamination. St Albans District council (SADC) commission intrusive investigations at
the SLC site. The bromate groundwater monitoring programme begins to establish the extent and
migration of the contamination.
2002 The monitoring network is expanded and special investigations of surface waters are commenced,
including the Ellenbrook and the River Colne. SLC is designated as ‘Contaminated Land’ based
upon the ‘significant pollutant linkage’ between bromate and bromide in the unsaturated zone
and groundwater in the underlying Chalk aquifer. The site is adopted as a ‘special site’ by the
Environment Agency (EA). The EA begin consultation and investigation as to the determination
of the ‘appropriate persons’.
2003–2004 Bromate concentrations continue to increase at Essendon PWS and the NNR wells, threat-
ening available water resources for pulic supply.
2005 Pumping trial is commenced at Hatfield PWS to assess the possibility of scavenge pumping to
protect down-gradient sources at Essendon PWS and the NNR wells. The EA issue a remediation
notice to two appropriate persons: Redland Minerals Ltd for the Bromate contamination, and Crest
Nicholson Residential Ltd for the Bromide contamination. Both parties appeal.
2007 Appeals to the remediation notice are heard by public inquiry. Scavenge pumping of the Hatfield
source is promoted as an interim remediation measure and is incorporated in the Inspectors Re-
mediation Notice. An abstraction license is granted by the EA to VWTV for up to 9 Ml/d for the
purposes of groundwater remediation only.
2009 Decision is reached by the Secretary of State to uphold a modified remediation notice against
Redland Minerals Ltd and against Crest Nicholson Residential Ltd. Both parties apply for judicial
review.
Appendix B
The bromate contamination plume represents a major threat to the long-term quality of a number of
strategic public water supply sources, and also many private supply sources. The main driver for the
research conducted so far, and the funding for the additional research, is the impact of the bromate
contamination on water quality and resource availability.
Bromate concentrations above the regulatory standard for drinking water of 10 µg l−1 bromate
(UK Water Supply (Water Quality) Regulations 2000) are exceeded, or close to exceedence in major
PWS sources operated by both VWP and TWUL. Elevated bromate concentrations in PWS sources are
resulting in significant cost (∼£2million since May 2000) to the Water Companies. These costs include:
• The loss of large 9ML/d source since May 2000 when Hatfield Bishops Rise was taken out of
supply;
• The cost of blending at Essendon. The future use of Essendon may be restricted depending on
future blending arrangements.
• The cost of extra treatment required from the NNR sources/River Lee at Hornsey WTW.
The impact on water resources has lead Veolia Water Partnership (VWP) to plan to install two al-
ternative sources outside of the plume area. These boreholes are likely to replace Hatfield and Essendon
by December 2008. The total cost of the relocation amounts to ∼£8 million. Additional bromate mitiga-
tion and treatment measures are being considered by TWUL to safeguard the continued use of the NNR
sources, which will entail further expenditure. The costs of these impacts of the bromate contamination
are being largely borne by the water companies (and ultimately the customers). Although the Contam-
inated Land regime being enforced by the Environment Agency may result in the identified ‘polluter’
being liable for costs, this process in likely to take many years and the outcome is uncertain.
Water companies are particularly concerned that the future movement of the plume may affect
additional large PWS sources. In order to evaluate the most appropriate strategies for the current and
future management of groundwater quality, it is essential that additional research focusing on bromate
behaviour and movement in the aquifer is undertaken. This will allow the water companies and the
Environment Agency to refine their understanding of the contamination within the aquifer and to make
predictions as to the future evolution of the plume.
For example, a better understanding of the nature of groundwater flow in the area, will help to
characterise and quantify the effect of pumping the Hatfield borehole has on concentrations at other key
locations (particularly Essendon and NNR sources). This will allow the feasibility of using scavenge
pumping as a means of controlling concentrations at these sources to be evaluated.
283
Appendix C
Analysis of Variance
Predictor Coef SE Coef T P
Source DF SS MS F P Constant 30.0122 0.5830 51.48 0.000
Regression 1 4555.7 4555.7 129.49 0.000 H(T-3) -1.4105 0.1160 -12.16 0.000
Residual Error 333 11715.2 35.2
Total 334 16270.9
S = 5.81719 R-Sq = 30.7% R-Sq(adj) = 30.5%
S = 5.76994 R-Sq = 31.9% R-Sq(adj) = 31.7% 335 cases used, 945 cases contain missing values
Analysis of Variance
Regression Analysis: Bromate as BrO3 (ug/l) versus H(T-2)
Source DF SS MS F P
The regression equation is Regression 1 4220.9 4220.9 116.64 0.000
Bromate as BrO3 (ug/l) = 30.5 - 1.58 H(T-2) Residual Error 333 12050.1 36.2
Total 334 16270.9
335 cases used, 945 cases contain missing values
Regression Analysis: Bromate as BrO3 (ug/l) versus H(T-5)
Predictor Coef SE Coef T P
Constant 30.4704 0.5382 56.62 0.000 The regression equation is
H(T-2) -1.5824 0.1104 -14.33 0.000 Bromate as BrO3 (ug/l) = 29.8 - 1.28 H(T-5)
S = 5.49730 R-Sq = 38.2% R-Sq(adj) = 38.0% 334 cases used, 946 cases contain missing values
Predictor Coef SE Coef T P Essendon Bromate V Hatfield Abstraction (T-2)
Constant 29.7661 0.6775 43.94 0.000
H(T-5) -1.2757 0.1308 -9.76 0.000 Normal Probability Plot Versus Fits
99.9
99 4
S = 6.11821 R-Sq = 22.3% R-Sq(adj) = 22.0%
90
2
Analysis of Variance 50
0
Percent
Source DF SS MS F P 10
-2
Regression 1 3562.6 3562.6 95.17 0.000 1
Residual Error 332 12427.6 37.4
Standardized Residual
0.1 -4
Total 333 15990.2 -4 -2 0 2 4 15 20 25 30
Standardized Residual Fitted Value
Regression Analysis: Bromate as BrO3 (ug/l) versus H(T-6) Histogram Versus Order
80
The regression equation is 4
Bromate as BrO3 (ug/l) = 28.7 - 1.09 H(T-6)
60
2
40
333 cases used, 947 cases contain missing values 0
Frequency
20 -2
Predictor Coef SE Coef T P 0 -4
Standardized Residual
Constant 28.6746 0.6490 44.19 0.000 -3.0 -1.5 0.0 1.5 3.0 4.5 1 00 00 00 00 00 00 00 00 00 00 00 00
H(T-6) -1.0912 0.1281 -8.52 0.000 Standardized Residual
1 2 3 4 5 6 7 8 9 10 11 12
Observation Order
S = 6.26975 R-Sq = 18.0% R-Sq(adj) = 17.7%
3
Regression Analysis: Bromate as BrO3 (ug/l) versus H(T-7)
2
The regression equation is
Bromate as BrO3 (ug/l) = 28.4 - 1.08 H(T-7) 1
0
333 cases used, 947 cases contain missing values
-1
Predictor Coef SE Coef T P
Standardized Residual
1
30
0
-1 20
Standardized Residual
Bromate as BrO3 (ug/l)
-2
10
-3
-4 0
0 20 40 60 80 100 120 140 0 1 2 3 4 5 6 7 8 9
SMD H(T-2)
SMD vs Date
140
Fitted Line Plot
120 Bromate as BrO3 (ug/l) = 30.47 - 1.582 H(T-2)
60 Regression
100 95% CI
95% PI
50
80 S 5.49730
R-Sq 38.2%
40 R-Sq(adj) 38.0%
60
40 30
20
0
10
01/01/2005 01/01/2006 01/01/2007 01/01/2008 01/01/2009
Date
0
0 1 2 3 4 5 6 7 8 9
H(T-2)
Regression Analysis: Bromate as BrO3 (ug/l) versus H(T-2), SMD - 6140
Analysis of Variance 0
Source DF SS MS F P -1
Standardized Residual
Regression 2 7979.4 3989.7 159.75 0.000
Residual Error 332 8291.5 25.0
Total 334 16270.9 -2
-3
Source DF Seq SS
H(T-2) 1 6207.5 01/01/2006 01/01/2007 01/01/2008 01/01/2009
SMD - 6140 1 1771.9
Date
Percent
10
-2
1
Standardized Residual
0.1 -4
-4 -2 0 2 4 15 20 25 30 35
Standardized Residual Fitted Value
4
60
45 2
30 0
Frequency
15 -2
-4
Standardized Residual
0
-3 -2 -1 0 1 2 3 4 1 00 00 00 00 00 00 00 00 00 00 00 00
1 2 3 4 5 6 7 8 9 10 11 12
Standardized Residual
Observation Order
Amwell Hill bromate versus Hatfield abstraction
S = 5.04850 R-Sq = 1.3% R-Sq(adj) = 0.6%
Regression Analysis: Bromate as BrO3 (ug/l) versus H(T)
Analysis of Variance 137 cases used, 1143 cases contain missing values
Source DF SS MS F P
Regression 1 66.10 66.10 2.59 0.110 Predictor Coef SE Coef T P
Residual Error 138 3518.36 25.50 Constant 9.6970 0.7308 13.27 0.000
Total 139 3584.46 H(T-3) -0.1923 0.1478 -1.30 0.196
Predictor Coef SE Coef T P Regression Analysis: Bromate as BrO3 (ug/l) versus H(T-5)
Constant 9.5796 0.6867 13.95 0.000
H(T-2) -0.1943 0.1462 -1.33 0.186 The regression equation is
Bromate as BrO3 (ug/l) = 10.6 - 0.332 H(T-5) Regression 1 27.61 27.61 1.16 0.283
Residual Error 126 2988.22 23.72
Total 127 3015.83
132 cases used, 1148 cases contain missing values
90 1
Regression Analysis: Bromate as BrO3 (ug/l) versus H(T-6)
50 0
Percent
The regression equation is
10 -1
Bromate as BrO3 (ug/l) = 10.5 - 0.311 H(T-6)
1
Standardized Residual
0.1 -2
130 cases used, 1150 cases contain missing values -4 -2 0 2 4 8 9 10 11
Standardized Residual Fitted Value
12 1
S = 4.86986 R-Sq = 3.2% R-Sq(adj) = 2.5%
8 0
Frequency
4 -1
Analysis of Variance
0 -2
Standardized Residual
Source DF SS MS F P
-1.50 -0.75 0.00 0.75 1.50 1 00 00 00 00 00 00 00 00 00 00 00 00
Regression 1 101.44 101.44 4.28 0.041 1 2 3 4 5 6 7 8 9 10 11 12
Standardized Residual
Residual Error 128 3035.59 23.72
Observation Order
Total 129 3137.03
Analysis of Variance
Source DF SS MS F P
Residuals Versus Date Amwell Hill Bromate V Hatfield Abstraction (T-5)
(response is Bromate as BrO3 (ug/l)) Bromate as BrO3 (ug/l) = 10.58 - 0.3325 H(T-5)
2 Regression
20 95% CI
95% PI
S 4.91246
1 15 R-Sq 3.6%
R-Sq(adj) 2.8%
10
0
5
-1
Standardized Residual
Bromate as BrO3 (ug/l)
0
-2
01/01/2006 01/01/2007 01/01/2008 01/01/2009 0 1 2 3 4 5 6 7 8 9
Date H(T-5)
Analysis of Variance
-1 Source DF SS MS F P
Standardized Residual
Regression 1 0.6864 0.6864 6.25 0.014
Residual Error 130 14.2844 0.1099
Total 131 14.9708
-2
0 20 40 60 80 100 120 140
6600 - Lee Chalk
log10(Amwell Hill Bromate) V Hatfield Abstraction (T-5)
Residuals Versus SMD
Normal Probability Plot Versus Fits (response is Log(BrO3))
99.9
99 1
90 1
0
50
-1
Percent
10
-2
1
0
Standardized Residual
0.1 -3
-4 -2 0 2 4 0.75 0.80 0.85 0.90 0.95
Standardized Residual Fitted Value
-1
Histogram Versus Order
20 1
Standardized Residual
15 0 -2
10 -1
Frequency
5 -2
-3
-3
Standardized Residual
0
-2.5 -2.0 -1.5 -1.0 -0.5 0.0 0.5 1.0 1 00 00 00 00 00 00 00 00 00 00 00 00 0 20 40 60 80 100 120 140
1 2 3 4 5 6 7 8 9 10 11 12
Standardized Residual 6600 - Lee Chalk
Observation Order
S 4.91246
15 R-Sq 3.6%
0 R-Sq(adj) 2.8%
10
-1
Standardized Residual
Bromate as BrO3 (ug/l)
-2
0
-3
01/01/2006 01/01/2007 01/01/2008 01/01/2009 0 1 2 3 4 5 6 7 8 9
Date H(T-5)
Source DF SS MS F P -1
Standardized Residual
Regression 2 8.2543 4.1272 79.27 0.000
Residual Error 129 6.7165 0.0521
Total 131 14.9708
-2
Source DF Seq SS
H(T-5) 1 0.6864 01/01/2006 01/01/2007 01/01/2008 01/01/2009
6600 - Lee Chalk 1 7.5679
Date
Percent
10 -1
1 -2
Standardized Residual
0.1
-4 -2 0 2 4 0.4 0.6 0.8 1.0 1.2
Standardized Residual Fitted Value
12 0
Frequency
-1
6
-2
Standardized Residual
0
-2.4 -1.6 -0.8 0.0 0.8 1.6 2.4 1 00 00 00 00 00 00 00 00 00 00 00 00
1 2 3 4 5 6 7 8 9 10 11 12
Standardized Residual
Observation Order
Hoddesdon bromate versus Hatfield abstraction S = 9.96880 R-Sq = 12.1% R-Sq(adj) = 11.6%
Analysis of Variance
Predictor Coef SE Coef T P
Constant 30.595 1.352 22.63 0.000
Source DF SS MS F P
H(T-3) -1.5435 0.2744 -5.63 0.000
Regression 1 2584.1 2584.1 26.05 0.000
Residual Error 161 15972.7 99.2
Total 162 18556.7
S = 9.63218 R-Sq = 16.8% R-Sq(adj) = 16.2%
Analysis of Variance
Predictor Coef SE Coef T P
Constant 32.913 1.401 23.49 0.000
Source DF SS MS F P
H(T-4) -1.8999 0.2673 -7.11 0.000
Regression 1 2084.5 2084.5 20.37 0.000
Residual Error 161 16472.2 102.3
Total 162 18556.7
S = 9.16736 R-Sq = 24.6% R-Sq(adj) = 24.1%
50
Analysis of Variance 0
Percent
10
Source DF SS MS F P 1 -2
Regression 1 3688.3 3688.3 43.29 0.000
Standardized Residual
0.1
Residual Error 153 13035.5 85.2 -4 -2 0 2 4 15 20 25 30 35
Total 154 16723.7 Standardized Residual Fitted Value
Frequency
-2
0
Standardized Residual
2
Regression Analysis: Bromate as BrO3 (ug/l) versus H(T-7)
0
151 cases used, 1129 cases contain missing values
-1
Standardized Residual
Source DF SS MS F P
Regression 1 3355.4 3355.4 39.04 0.000
Residual Error 149 12805.7 85.9
Total 150 16161.1
The regression equation is
Residuals Versus SMD Log(BrO3) = 1.48 - 0.0288 H(T-4)
(response is Bromate as BrO3 (ug/l))
4 157 cases used, 1123 cases contain missing values
Standardized Residual
Total 156 4.73220
-2
Residual Plots for Log(BrO3)
-3
0 20 40 60 80 100 120 140 Normal Probability Plot Versus Fits
99.9 3.0
6600 - Lee Chalk 99
90 1.5
Percent
Bromate as BrO3 (ug/l) = 32.91 - 1.900 H(T-4) 10
-1.5
Regression 1
Standardized Residual
60 0.1 -3.0
95% CI
-4 -2 0 2 4 1.25 1.30 1.35 1.40 1.45
95% PI Standardized Residual Fitted Value
50 S 9.16736
R-Sq 24.6% Histogram Versus Order
40 R-Sq(adj) 24.1% 30 3.0
1.5
20
30 0.0
10
Frequency
-1.5
20
0 -3.0
Standardized Residual
-3 -2 -1 0 1 2 1 00 00 00 00 00 00 00 00 00 00 00 00
0
Regression Analysis: Bromate as BrO3 versus H(T-4), 6600 - Lee Chalk
0 1 2 3 4 5 6 7 8 9 The regression equation is
H(T-4) Bromate as BrO3 (ug/l) = 31.1 - 2.19 H(T-4) + 0.0602 6600 - Lee Chalk
3
Analysis of Variance
Source DF SS MS F P 2
Regression 2 5208.3 2604.1 33.24 0.000
Residual Error 154 12063.5 78.3
Total 156 17271.8
1
Source DF Seq SS
H(T-4) 1 4245.5 0
6600 - Lee Chalk 1 962.7
-1
Standardized Residual
Hoddesdon Bromate V Hatfield Abstraction (T-4), SMD-Lee
Normal Probability Plot Versus Fits -2
99.9
3.0
99
-3
90 1.5
01/01/2006 01/01/2007 01/01/2008 01/01/2009
50 0.0 Date
Percent
10
-1.5
1
Standardized Residual
0.1 -3.0 Residuals Versus SMD
-4 -2 0 2 4 15 20 25 30 35
Standardized Residual Fitted Value
(response is Bromate as BrO3 (ug/l))
30 1.5 2
20 0.0
1
Frequency
10 -1.5
0 -3.0
Standardized Residual
-2 -1 0 1 2 3 1 00 00 00 00 00 00 00 00 00 00 00 00 0
1 2 3 4 5 6 7 8 9 10 11 12
Standardized Residual
Observation Order
-1
Standardized Residual
-2
-3
0 20 40 60 80 100 120 140
6600 - Lee Chalk
Middlefield Road bromate versus Hatfield abstraction S = 7.86399 R-Sq = 4.1% R-Sq(adj) = 3.5%
Regression Analysis: Bromate as BrO3 (ug/l) versus H(T)
Analysis of Variance
The regression equation is
Bromate as BrO3 (ug/l) = 20.1 - 0.599 H(T) Source DF SS MS F P
Regression 1 430.95 430.95 6.97 0.009
Residual Error 164 10142.15 61.84
167 cases used, 1113 cases contain missing values Total 165 10573.10
Analysis of Variance 165 cases used, 1115 cases contain missing values
Source DF SS MS F P
Regression 1 433.50 433.50 7.02 0.009 Predictor Coef SE Coef T P
Residual Error 165 10191.39 61.77 Constant 19.407 1.090 17.80 0.000
Total 166 10624.89 H(T-3) -0.4167 0.2211 -1.88 0.061
Analysis of Variance 163 cases used, 1117 cases contain missing values
Source DF SS MS F P
Regression 1 481.60 481.60 7.83 0.006 Predictor Coef SE Coef T P
Residual Error 165 10143.29 61.47 Constant 20.879 1.156 18.05 0.000
Total 166 10624.89 H(T-4) -0.6717 0.2178 -3.08 0.002
Source DF SS MS F P
Predictor Coef SE Coef T P
Regression 1 768.65 768.65 12.84 0.000
Constant 21.110 1.034 20.41 0.000
Residual Error 159 9518.54 59.87
H(T-8) -0.8979 0.2174 -4.13 0.000
Total 160 10287.19
Source DF SS MS F P
Predictor Coef SE Coef T P
Regression 1 659.74 659.74 10.87 0.001
Constant 20.4131 0.9986 20.44 0.000
Residual Error 157 9531.39 60.71
H(T-9) -0.7342 0.2090 -3.51 0.001
Total 158 10191.13
Analysis of Variance
Middlefield Rd Bromate V Hatfield Abstraction (T-8)
Residuals Versus SMD
Normal Probability Plot Versus Fits (response is Bromate as BrO3 (ug/l))
99.9 4
99 4
90 2
50
3
0
Percent
10
1
2
-2
Standardized Residual
0.1
-4 -2 0 2 4 14 16 18 20 22
Standardized Residual Fitted Value 1
Standardized Residual
2
30 -1
20 0
Frequency
10 -2
-2
0
Standardized Residual
-2 -1 0 1 2 3 1 00 00 00 00 00 00 00 00 00 00 00 00 0 20 40 60 80 100 120 140
1 2 3 4 5 6 7 8 9 10 11 12
Standardized Residual 6600 - Lee Chalk
Observation Order
1
20
0
10
Standardized Residual
Bromate as BrO3 (ug/l)
-1
-2 0
1
Predictor Coef SE Coef T P
Constant 1.28012 0.02699 47.43 0.000 0
H(T-8) -0.020048 0.005675 -3.53 0.001
-1
S = 0.199886 R-Sq = 7.1% R-Sq(adj) = 6.5%
-2
Analysis of Variance -3
Source DF SS MS F P
Standardized Residual
Regression 1 0.49866 0.49866 12.48 0.001 -4
Residual Error 164 6.55252 0.03995
Total 165 7.05119 -5
-6
Log10(Middlefield Rd Bromate) V Hatfield Abstraction (T-8)
01/01/2006 01/01/2007 01/01/2008 01/01/2009
Normal Probability Plot Versus Fits Date
99.9
2
99
90 0
Residuals Versus SMD
50 -2 (response is Log(BrO3))
Percent
10 -4 3
1
Standardized Residual
0.1
-6 2
-6 -4 -2 0 2 1.10 1.15 1.20 1.25 1.30
Standardized Residual Fitted Value 1
Histogram Versus Order 0
40 2
-1
30 0
20 -2
-2
Frequency
10 -4 -3
-6
Standardized Residual
0
Standardized Residual
-4
-6.0 -4.5 -3.0 -1.5 0.0 1.5 1 00 00 00 00 00 00 00 00 00 00 00 00
1 2 3 4 5 6 7 8 9 10 11 12
Standardized Residual
Observation Order
-5
-6
0 20 40 60 80 100 120 140
6600 - Lee Chalk
412 4.39 0.0273 1.1694 0.0300 -1.1421 -5.78R
432 0.01 1.6854 1.2564 0.0377 0.4289 2.18R
Log10(Middlefield Rd Bromate) V Hatfield Abstraction (T-8) 860 7.80 0.4624 1.1209 0.0273 -0.6585 -3.32R
Log(BrO3) = 1.280 - 0.02005 H(T-8)
R denotes an observation with a large standardized residual.
1.8 Regression
95% CI
1.6 Residual Plots for Log(BrO3)
95% PI
Normal Probability Plot Versus Fits
1.4 S 0.199886
99.9
R-Sq 7.1% 2
99
1.2 R-Sq(adj) 6.5%
90 0
1.0 50 -2
Percent
0.8 10
-4
Log(BrO3)
1
0.6
Standardized Residual
0.1 -6
-6 -4 -2 0 2 1.10 1.15 1.20 1.25 1.30
0.4 Standardized Residual Fitted Value
Frequency
10 -4
0
-6.0 -4.5 -3.0 -1.5 0.0 1.5 1 00 00 00 00 00 00 00 00 00 00 00 00
1 2 3 4 5 6 7 8 9 10 11 12
The regression equation is Standardized Residual
Log(BrO3) = 1.29 - 0.0191 H(T-8) - 0.000323 6600 - Lee Chalk Observation Order
Analysis of Variance
Source DF SS MS F P
Regression 2 0.53042 0.26521 6.63 0.002
Residual Error 163 6.52077 0.04000
Total 165 7.05119
Source DF Seq SS
H(T-8) 1 0.49866
6600 - Lee Chalk 1 0.03176
Unusual Observations
-1
-2
-3
Standardized Residual
-4
-5
-6
01/01/2006 01/01/2007 01/01/2008 01/01/2009
Date
Rye Common versus Hatfield abstraction S = 6.64265 R-Sq = 0.3% R-Sq(adj) = 0.0%
Analysis of Variance
Predictor Coef SE Coef T P
Constant 9.5952 0.9208 10.42 0.000
Source DF SS MS F P
H(T-3) -0.0919 0.1898 -0.48 0.629
Regression 1 4.60 4.60 0.10 0.747
Residual Error 160 7079.26 44.25
Total 161 7083.87
S = 6.64902 R-Sq = 0.1% R-Sq(adj) = 0.0%
Analysis of Variance
Predictor Coef SE Coef T P
Constant 10.1653 0.9752 10.42 0.000
Source DF SS MS F P
H(T-4) -0.2124 0.1869 -1.14 0.257
Regression 1 40.99 40.99 0.93 0.336
Residual Error 160 7042.88 44.02
Total 161 7083.87
S = 6.62719 R-Sq = 0.8% R-Sq(adj) = 0.2%
Source DF SS MS F P
Predictor Coef SE Coef T P
Regression 1 101.31 101.31 2.31 0.130
Constant 10.7462 0.9074 11.84 0.000
Residual Error 159 6967.44 43.82
H(T-8) -0.3884 0.1935 -2.01 0.046
Total 160 7068.74
Source DF SS MS F P
Predictor Coef SE Coef T P
Regression 1 88.57 88.57 2.02 0.157
Constant 10.1108 0.8765 11.54 0.000
Residual Error 159 6980.18 43.90
H(T-9) -0.2256 0.1848 -1.22 0.224
Total 160 7068.74
Analysis of Variance
Source DF SS MS F P
Regression 1 225.39 225.39 5.24 0.023
Rye Common Bromate V Hatfield Abstraction
Residuals Versus SMD-6600
Normal Probability Plot Versus Fits (response is Bromate as BrO3 (ug/l))
99.9
99 3 4
90 2
50 1 3
Percent
10 0
1 -1 2
Standardized Residual
0.1
-4 -2 0 2 4 7 8 9 10 11
Standardized Residual Fitted Value
1
Histogram Versus Order
16 0
3
12 2
Standardized Residual
8 1 -1
0
Frequency
4
-1
-2
Standardized Residual
0
-1.50 -0.75 0.00 0.75 1.50 2.25 3.00 1 0 0 0 0 0 0 0 0 0 0 0 0 0 20 40 60 80 100 120 140
10 20 30 40 50 60 70 80 90 100 110 120
Standardized Residual 6600 - Lee Chalk
Observation Order
0 Analysis of Variance
Standardized Residual
Source DF SS MS F P
-1 Regression 1 0.1769 0.1769 1.07 0.303
Residual Error 159 26.3312 0.1656
Total 160 26.5081
-2
01/01/2006 01/01/2007 01/01/2008 01/01/2009
Date
Log10(Rye Common Bromate) V Hatfield Abstraction
Normal Probability Plot Versus Fits
99.9 2 Regression Analysis: Bromate as BrO3 versus H(T-7), 6600 - Lee Chalk
99
1
90 The regression equation is
0
50 Bromate as BrO3 (ug/l) = 11.8 - 0.350 H(T-7) - 0.0206 6600 - Lee Chalk
-1
Percent
10
-2
1
161 cases used, 1119 cases contain missing values
Standardized Residual
0.1
-4 -2 0 2 4 0.750 0.775 0.800 0.825 0.850
Standardized Residual Fitted Value
Predictor Coef SE Coef T P
Histogram Versus Order
Constant 11.786 1.034 11.40 0.000
2
20
H(T-7) -0.3499 0.1961 -1.78 0.076
1 6600 - Lee Chalk -0.02063 0.01239 -1.66 0.098
15
0
10 -1
Frequency
5 -2 S = 6.52423 R-Sq = 4.9% R-Sq(adj) = 3.7%
0
Standardized Residual
-2.25 -1.50 -0.75 0.00 0.75 1.50 1 0 0 0 0 0 0 0 0 0 0 0 0
10 20 30 40 50 60 70 80 90 100 110 120
Standardized Residual
Observation Order
Analysis of Variance
Source DF SS MS F P
Residuals Versus Date Regression 2 343.38 171.69 4.03 0.020
(response is Log(BrO3)) Residual Error 158 6725.37 42.57
2 Total 160 7068.74
1
Source DF Seq SS
H(T-7) 1 225.39
6600 - Lee Chalk 1 117.98
0
Unusual Observations
-1
Bromate
as BrO3
Standardized Residual
Obs H(T-7) (ug/l) Fit SE Fit Residual St Resid
-2
278 0.00 27.300 11.699 1.012 15.601 2.42R
286 0.41 33.200 11.554 0.963 21.646 3.35R
307 0.01 25.406 11.210 0.936 14.196 2.20R
-3
01/01/2006 01/01/2007 01/01/2008 01/01/2009 R denotes an observation with a large standardized residual.
Date
50 1
Percent
1 0
10
1
-1
Standardized Residual
0.1
0 -4 -2 0 2 4 5.0 7.5 10.0 12.5
Standardized Residual Fitted Value
Standardized Residual
-2 20 1
0
Frequency
10
-1
0
Standardized Residual
-3
-1.6 -0.8 0.0 0.8 1.6 2.4 3.2 1 00 00 00 00 00 00 00 00 00 00 00 00
0 20 40 60 80 100 120 140 Standardized Residual
1 2 3 4 5 6 7 8 9 10 11 12
6600 - Lee Chalk Observation Order
Turnford bromate versus Hatfield abstraction
Residuals Versus Date Regression Analysis: Bromate as BrO3 (ug/l) versus H(T)
(response is Bromate as BrO3 (ug/l))
4 The regression equation is
Bromate as BrO3 (ug/l) = 21.0 - 1.04 H(T)
3
Standardized Residual
S = 8.92022 R-Sq = 9.3% R-Sq(adj) = 8.7%
-1
-2 Analysis of Variance
01/01/2006 01/01/2007 01/01/2008 01/01/2009
Date Source DF SS MS F P
Regression 1 1315.0 1315.0 16.53 0.000
Residual Error 162 12890.4 79.6
Residuals Versus SMD-6600 Total 163 14205.4
(response is Bromate as BrO3 (ug/l))
4
Unusual Observations
3 Bromate
as BrO3
Obs H(T) (ug/l) Fit SE Fit Residual St Resid
2 5 0.03 40.813 21.010 1.272 19.803 2.24R
12 0.23 39.123 20.801 1.230 18.322 2.07R
1 286 0.00 43.500 21.041 1.279 22.459 2.54R
300 0.01 40.816 21.030 1.277 19.786 2.24R
307 1.84 45.790 19.121 0.919 26.669 3.01R
0 319 2.99 37.711 17.920 0.760 19.791 2.23R
495 0.00 0.600 21.038 1.278 -20.438 -2.32R
Standardized Residual
-1 1146 4.80 34.400 16.035 0.714 18.365 2.07R
Analysis of Variance
Source DF SS MS F P
Regression 1 1219.1 1219.1 15.21 0.000
Residual Error 162 12986.2 80.2
Total 163 14205.4
Unusual Observations 164 cases used, 1116 cases contain missing values
Bromate
as BrO3 Predictor Coef SE Coef T P
Obs H(T-1) (ug/l) Fit SE Fit Residual St Resid Constant 20.172 1.184 17.04 0.000
5 0.00 40.813 20.435 1.190 20.378 2.30R H(T-3) -0.8819 0.2407 -3.66 0.000
12 0.44 39.123 20.005 1.103 19.118 2.15R
271 0.65 38.100 19.801 1.062 18.299 2.06R
286 0.00 43.500 20.435 1.190 23.065 2.60R S = 8.99859 R-Sq = 7.7% R-Sq(adj) = 7.1%
300 0.37 40.816 20.074 1.116 20.742 2.33R
307 1.98 45.790 18.504 0.841 27.286 3.06R
319 3.00 37.711 17.509 0.731 20.202 2.26R Analysis of Variance
495 0.00 0.600 20.431 1.189 -19.831 -2.23R
1146 4.80 34.400 15.758 0.738 18.642 2.09R Source DF SS MS F P
Regression 1 1087.5 1087.5 13.43 0.000
R denotes an observation with a large standardized residual. Residual Error 162 13117.9 81.0
Total 163 14205.4
Unusual Observations 164 cases used, 1116 cases contain missing values
Bromate
as BrO3 Predictor Coef SE Coef T P
Obs H(T-2) (ug/l) Fit SE Fit Residual St Resid Constant 22.115 1.265 17.49 0.000
5 0.00 40.813 19.849 1.138 20.964 2.34R H(T-4) -1.2359 0.2426 -5.09 0.000
12 0.01 39.123 19.840 1.136 19.283 2.15R
271 0.00 38.100 19.849 1.138 18.251 2.04R
286 0.00 43.500 19.849 1.138 23.651 2.64R S = 8.69378 R-Sq = 13.8% R-Sq(adj) = 13.3%
300 0.01 40.816 19.840 1.136 20.976 2.34R
307 0.01 45.790 19.840 1.136 25.950 2.90R
319 3.00 37.711 17.247 0.722 20.464 2.28R Analysis of Variance
495 0.00 0.600 19.846 1.138 -19.246 -2.15R
1146 4.79 34.400 15.691 0.758 18.709 2.08R Source DF SS MS F P
Regression 1 1961.1 1961.1 25.95 0.000
R denotes an observation with a large standardized residual. Residual Error 162 12244.2 75.6
Total 163 14205.4
Unusual Observations
Regression Analysis: Bromate as BrO3 (ug/l) versus H(T-5)
Bromate
as BrO3
The regression equation is
Obs H(T-6) (ug/l) Fit SE Fit Residual St Resid
Bromate as BrO3 (ug/l) = 21.7 - 1.19 H(T-5)
12 0.02 39.123 20.827 1.256 18.296 2.11R
286 0.00 43.500 20.847 1.260 22.653 2.61R
300 3.00 40.816 17.784 0.751 23.032 2.64R
163 cases used, 1117 cases contain missing values
307 0.01 45.790 20.837 1.258 24.953 2.88R
319 2.99 37.711 17.794 0.752 19.917 2.28R
495 0.01 0.600 20.841 1.259 -20.241 -2.34R
Predictor Coef SE Coef T P
Constant 21.684 1.245 17.42 0.000
R denotes an observation with a large standardized residual.
H(T-5) -1.1898 0.2420 -4.92 0.000
S = 8.57627 R-Sq = 13.1% R-Sq(adj) = 12.5% Regression Analysis: Bromate as BrO3 (ug/l) versus H(T-7)
Source DF SS MS F P
Regression 1 1777.5 1777.5 24.17 0.000 163 cases used, 1117 cases contain missing values
Residual Error 161 11841.9 73.6
Total 162 13619.4
Predictor Coef SE Coef T P
Constant 21.315 1.261 16.90 0.000
Unusual Observations H(T-7) -1.1460 0.2547 -4.50 0.000
Bromate
as BrO3 S = 8.66839 R-Sq = 11.2% R-Sq(adj) = 10.6%
Obs H(T-5) (ug/l) Fit SE Fit Residual St Resid
12 0.02 39.123 21.660 1.241 17.463 2.06R
286 0.00 43.500 21.684 1.245 21.816 2.57R Analysis of Variance
300 1.09 40.816 20.387 1.032 20.429 2.40R
307 0.01 45.790 21.672 1.243 24.118 2.84R Source DF SS MS F P
319 2.99 37.711 18.127 0.746 19.584 2.29R Regression 1 1521.7 1521.7 20.25 0.000
495 0.00 0.600 21.679 1.244 -21.079 -2.48R Residual Error 161 12097.7 75.1
1133 4.92 33.300 15.826 0.687 17.474 2.04R Total 162 13619.4
Bromate
The regression equation is
as BrO3
Bromate as BrO3 (ug/l) = 20.3 - 0.978 H(T-8)
Obs H(T-9) (ug/l) Fit SE Fit Residual St Resid
12 0.00 39.123 20.243 1.142 18.880 2.17R
229 1.99 36.686 18.330 0.816 18.356 2.10R
163 cases used, 1117 cases contain missing values
271 0.00 38.100 20.243 1.142 17.857 2.06R
286 0.00 43.500 20.243 1.142 23.257 2.68R
300 0.01 40.816 20.233 1.140 20.583 2.37R
Predictor Coef SE Coef T P
307 0.01 45.790 20.233 1.140 25.557 2.94R
Constant 20.296 1.176 17.26 0.000
319 3.00 37.711 17.360 0.716 20.351 2.33R
H(T-8) -0.9783 0.2479 -3.95 0.000
495 0.01 0.600 20.235 1.140 -19.635 -2.26R
1139 5.04 33.800 15.402 0.740 18.398 2.11R
S = 8.78231 R-Sq = 8.8% R-Sq(adj) = 8.3%
R denotes an observation with a large standardized residual.
Analysis of Variance
Turnford Bromate V Hatfield Abstraction
Source DF SS MS F P
Normal Probability Plot Versus Fits
Regression 1 1201.6 1201.6 15.58 0.000
99.9 3.0
Residual Error 161 12417.8 77.1 99
Total 162 13619.4 90 1.5
50 0.0
Percent
Unusual Observations 10
-1.5
1
Standardized Residual
Frequency
10 -1.5
307 0.37 45.790 19.934 1.103 25.856 2.97R
0 -3.0
Standardized Residual
Regression Analysis: Bromate as BrO3 (ug/l) versus H(T-9) Residuals Versus Date
(response is Bromate as BrO3 (ug/l))
The regression equation is 3
Bromate as BrO3 (ug/l) = 20.2 - 0.961 H(T-9)
2
163 cases used, 1117 cases contain missing values
1
Analysis of Variance
-3
Source DF SS MS F P 01/01/2006 01/01/2007 01/01/2008 01/01/2009
Regression 1 1267.9 1267.9 16.53 0.000 Date
Residual Error 161 12351.5 76.7
Total 162 13619.4
Constant 22.413 1.374 16.31 0.000
Residuals Versus SMD-6600 H(T-4) -1.1862 0.2588 -4.58 0.000
(response is Bromate as BrO3 (ug/l)) 6600 - Lee Chalk -0.00937 0.01673 -0.56 0.576
3
S = 8.71226 R-Sq = 14.0% R-Sq(adj) = 12.9%
2
Analysis of Variance
1
Source DF SS MS F P
0 Regression 2 1984.89 992.45 13.08 0.000
Residual Error 161 12220.46 75.90
Total 163 14205.35
-1
Standardized Residual
-2 Source DF Seq SS
H(T-4) 1 1961.10
6600 - Lee Chalk 1 23.79
-3
0 20 40 60 80 100 120 140
6600 - Lee Chalk
Turnford Bromate V Hatfield Abstraction & SMD-Lee
Normal Probability Plot Versus Fits
Regression Analysis: Log(BrO3) versus H(T-4) 99.9 3.0
99
90 1.5
The regression equation is
Log(BrO3) = 1.26 - 0.0244 H(T-4) 50 0.0
Percent
10
-1.5
1
Standardized Residual
164 cases used, 1116 cases contain missing values 0.1 -3.0
-4 -2 0 2 4 10 15 20
Standardized Residual Fitted Value
20 0.0
S = 0.249658 R-Sq = 7.1% R-Sq(adj) = 6.5%
Frequency
10 -1.5
0 -3.0
Standardized Residual
-2 -1 0 1 2 1 0 0 0 0 0 0 0 0 0 0 0 0
Analysis of Variance Standardized Residual
10 20 30 40 50 60 70 80 90 100 110 120
Observation Order
Source DF SS MS F P
Regression 1 0.76735 0.76735 12.31 0.001
Residual Error 162 10.09733 0.06233 Residuals Versus Date
Total 163 10.86468 (response is Bromate as BrO3 (ug/l))
3
Unusual Observations
2
Obs H(T-4) Log(BrO3) Fit SE Fit Residual St Resid
495 0.00 -0.2218 1.2613 0.0363 -1.4831 -6.00R 1
530 0.00 0.6049 1.2613 0.0363 -0.6564 -2.66R
-3
164 cases used, 1116 cases contain missing values 01/01/2006 01/01/2007 01/01/2008 01/01/2009
Date
-1
Predictor Coef SE Coef T P
Standardized Residual
Constant 25.024 1.186 21.10 0.000
-2 H(T) -1.4599 0.2415 -6.04 0.000
-3
S = 7.59608 R-Sq = 20.8% R-Sq(adj) = 20.2%
0 20 40 60 80 100 120 140
6600 - Lee Chalk
Analysis of Variance
Source DF SS MS F P
Turnford Bromate V Hatfield Abstraction Regression 1 2107.6 2107.6 36.53 0.000
Bromate as BrO3 (ug/l) = 22.12 - 1.236 H(T-4) Residual Error 139 8020.4 57.7
50 Total 140 10128.0
Regression
95% C I
95% PI
40 Unusual Observations
S 8.69378
R-Sq 13.8%
R-Sq(adj) 13.3% Bromate
30
as BrO3
Obs H(T) (ug/l) Fit SE Fit Residual St Resid
20 12 0.23 40.651 24.689 1.139 15.962 2.13R
19 0.20 41.973 24.732 1.145 17.241 2.30R
286 0.00 42.700 25.024 1.186 17.676 2.36R
10 307 1.84 47.352 22.338 0.846 25.014 3.31R
313 2.99 36.112 20.659 0.697 15.453 2.04R
Analysis of Variance
Source DF SS MS F P
Regression 1 1440.2 1440.2 23.04 0.000
Residual Error 139 8687.8 62.5
Total 140 10128.0 Regression Analysis: Bromate as BrO3 (ug/l) versus H(T-3)
Bromate
The regression equation is
as BrO3
Bromate as BrO3 (ug/l) = 25.6 - 1.53 H(T-5)
Obs H(T-6) (ug/l) Fit SE Fit Residual St Resid
12 0.02 40.651 24.695 1.153 15.956 2.16R
19 0.06 41.973 24.639 1.146 17.334 2.35R
140 cases used, 1140 cases contain missing values
286 0.00 42.700 24.723 1.157 17.977 2.43R
300 3.00 39.185 20.523 0.689 18.662 2.51R
307 0.01 47.352 24.709 1.155 22.643 3.07R
Predictor Coef SE Coef T P
400 3.60 0.897 19.683 0.646 -18.786 -2.52R
Constant 25.570 1.139 22.45 0.000
522 0.06 9.200 24.636 1.145 -15.436 -2.09R
H(T-5) -1.5311 0.2191 -6.99 0.000
R denotes an observation with a large standardized residual.
S = 7.22703 R-Sq = 26.1% R-Sq(adj) = 25.6%
Unusual Observations
Regression Analysis: Bromate as BrO3 (ug/l) versus H(T-6)
Bromate
as BrO3 H(T-9) -1.2799 0.2238 -5.72 0.000
Obs H(T-7) (ug/l) Fit SE Fit Residual St Resid
12 0.03 40.651 24.093 1.177 16.558 2.19R
19 0.23 41.973 23.833 1.137 18.140 2.40R S = 7.56065 R-Sq = 19.2% R-Sq(adj) = 18.6%
271 0.20 39.300 23.872 1.143 15.428 2.04R
286 0.41 42.700 23.599 1.101 19.101 2.52R
300 3.00 39.185 20.230 0.697 18.955 2.49R Analysis of Variance
307 0.01 47.352 24.119 1.181 23.233 3.07R
400 5.41 0.897 17.094 0.726 -16.197 -2.12R Source DF SS MS F P
558 0.00 7.800 24.130 1.183 -16.330 -2.16R Regression 1 1869.2 1869.2 32.70 0.000
Residual Error 138 7888.6 57.2
R denotes an observation with a large standardized residual. Total 139 9757.8
Bromate
The regression equation is
as BrO3
Bromate as BrO3 (ug/l) = 23.6 - 1.25 H(T-8)
Obs H(T-9) (ug/l) Fit SE Fit Residual St Resid
12 0.00 40.651 23.780 1.073 16.871 2.25R
19 0.01 41.973 23.768 1.071 18.205 2.43R
140 cases used, 1140 cases contain missing values
271 0.00 39.300 23.780 1.073 15.520 2.07R
286 0.00 42.700 23.780 1.073 18.920 2.53R
300 0.01 39.185 23.768 1.071 15.417 2.06R
Predictor Coef SE Coef T P
307 0.01 47.352 23.768 1.071 23.584 3.15R
Constant 23.634 1.119 21.12 0.000
400 3.88 0.897 18.815 0.639 -17.918 -2.38R
H(T-8) -1.2457 0.2375 -5.24 0.000
403 3.60 0.665 19.173 0.641 -18.508 -2.46R
445 0.01 8.547 23.768 1.071 -15.221 -2.03R
538 0.00 8.600 23.777 1.072 -15.177 -2.03R
S = 7.67845 R-Sq = 16.6% R-Sq(adj) = 16.0%
558 0.00 7.800 23.777 1.072 -15.977 -2.13R
Source DF SS MS F P
Regression 1 1621.5 1621.5 27.50 0.000 Turnford Bromate V Hatfield Abstraction
Residual Error 138 8136.3 59.0
Normal Probability Plot Versus Fits
Total 139 9757.8
99.9
3.0
99
90 1.5
Unusual Observations
50 0.0
Percent
Bromate 10
-1.5
as BrO3 1
Standardized Residual
10 -1.5
0 -3.0
Standardized Residual
1
Predictor Coef SE Coef T P
Constant 24.997 1.209 20.67 0.000
0 H(T-5) -1.6493 0.2347 -7.03 0.000
6600 - Lee Chalk 0.02030 0.01473 1.38 0.170
-1
Standardized Residual
S = 7.20359 R-Sq = 27.1% R-Sq(adj) = 26.1%
-2
-3 Analysis of Variance
01/01/2006 01/01/2007 01/01/2008 01/01/2009
Date
Source DF SS MS F P
Regression 2 2648.6 1324.3 25.52 0.000
Residual Error 137 7109.2 51.9
Total 139 9757.8
Residuals Versus SMD
(response is Bromate as BrO3 (ug/l))
3
Source DF Seq SS
H(T-5) 1 2550.1
6600 - Lee Chalk 1 98.6
2
1 Unusual Observations
Bromate
0 as BrO3
Obs H(T-5) (ug/l) Fit SE Fit Residual St Resid
-1 19 0.17 41.973 26.982 1.641 14.991 2.14R
98 0.00 * 27.670 1.902 * * X
Standardized Residual
99 0.00 * 27.695 1.916 * * X
-2 108 0.00 * 27.616 1.871 * * X
124 0.00 11.098 26.741 1.419 -15.643 -2.21R
-3 286 0.00 42.700 25.084 1.189 17.616 2.48R
0 20 40 60 80 100 120 140 300 1.09 39.185 23.694 0.955 15.491 2.17R
6600 - Lee Chalk 307 0.01 47.352 25.544 1.134 21.808 3.07R
400 6.96 0.897 16.024 1.155 -15.127 -2.13R
403 6.96 0.665 16.064 1.176 -15.399 -2.17R
Turnford Bromate V Hatfield Abstraction 417 4.43 34.924 19.858 0.987 15.066 2.11R
Bromate as BrO3 (ug/l) = 25.57 - 1.531 H(T-5) 530 1.33 5.911 22.797 1.047 -16.886 -2.37R
50 Regression R denotes an observation with a large standardized residual.
95% C I
X denotes an observation whose X value gives it large leverage.
95% PI
40
S 7.22703
R-Sq 26.1%
R-Sq(adj) 25.6%
30
20
10
0 1 2 3 4 5 6 7 8 9
H(T-5)
Turnford Bromate V Hatfield Abstraction & SMD-lee
Amwell Marsh bromate versus Hatfield abstraction
Normal Probability Plot Versus Fits
99.9
Regression Analysis: Bromate as BrO3 (ug/l) versus H(T)
3.0
99
Percent
10
-1.5
1 185 cases used, 1095 cases contain missing values
Standardized Residual
0.1 -3.0
-4 -2 0 2 4 10 15 20 25
Standardized Residual Fitted Value
Predictor Coef SE Coef T P
Histogram Versus Order Constant 13.9156 0.5481 25.39 0.000
3.0 H(T) -0.5823 0.1147 -5.08 0.000
30
1.5
20
0.0
S = 4.12680 R-Sq = 12.3% R-Sq(adj) = 11.9%
Frequency
10 -1.5
-3.0
Standardized Residual
0
-2 -1 0 1 2 3 1 0 0 0 0 0 0 0 0 0 0 0 0 Analysis of Variance
10 20 30 40 50 60 70 80 90 100 110 120
Standardized Residual
Observation Order
Source DF SS MS F P
Regression 1 438.71 438.71 25.76 0.000
Residuals Versus Date Residual Error 183 3116.57 17.03
(response is Bromate as BrO3 (ug/l)) Total 184 3555.28
0
185 cases used, 1095 cases contain missing values
-1
Standardized Residual
Predictor Coef SE Coef T P
-2 Constant 13.5558 0.5222 25.96 0.000
H(T-1) -0.5233 0.1131 -4.63 0.000
-3
01/01/2006 01/01/2007 01/01/2008 01/01/2009
Date
S = 4.17033 R-Sq = 10.5% R-Sq(adj) = 10.0%
Analysis of Variance
Residuals Versus SMD
(response is Bromate as BrO3 (ug/l))
Source DF SS MS F P
Regression 1 372.61 372.61 21.42 0.000
3
Residual Error 183 3182.67 17.39
Total 184 3555.28
2
1
Regression Analysis: Bromate as BrO3 (ug/l) versus H(T-2)
0
The regression equation is
Bromate as BrO3 (ug/l) = 13.2 - 0.467 H(T-2)
-1
Standardized Residual
-2 185 cases used, 1095 cases contain missing values
Analysis of Variance 184 cases used, 1096 cases contain missing values
Source DF SS MS F P
Regression 1 318.43 318.43 18.00 0.000 Predictor Coef SE Coef T P
Residual Error 183 3236.86 17.69 Constant 14.7941 0.5351 27.65 0.000
Total 184 3555.28 H(T-5) -0.7611 0.1073 -7.09 0.000
Analysis of Variance 184 cases used, 1096 cases contain missing values
Source DF SS MS F P
Regression 1 490.86 490.86 29.31 0.000 Predictor Coef SE Coef T P
Residual Error 183 3064.42 16.75 Constant 14.3209 0.5288 27.08 0.000
Total 184 3555.28 H(T-6) -0.6804 0.1095 -6.22 0.000
Analysis of Variance 184 cases used, 1096 cases contain missing values
Source DF SS MS F P
Regression 1 759.18 759.18 49.69 0.000 Predictor Coef SE Coef T P
Residual Error 183 2796.10 15.28 Constant 14.2658 0.5244 27.21 0.000
Total 184 3555.28 H(T-7) -0.6859 0.1112 -6.17 0.000
Percent
10
-2
1
Standardized Residual
0.1
Regression Analysis: Bromate as BrO3 (ug/l) versus H(T-8) -4 -2 0 2 4 8 10 12 14
Standardized Residual Fitted Value
15 0
184 cases used, 1096 cases contain missing values 10
Frequency
5 -2
Standardized Residual
Predictor Coef SE Coef T P 0
-3.00 -2.25 -1.50 -0.75 0.00 0.75 1.50 1 0 0 0 0 0 0 0 0 0 0 0 0
Constant 13.8845 0.5215 26.62 0.000 Standardized Residual
10 20 30 40 50 60 70 80 90 100 110 120
H(T-8) -0.6197 0.1151 -5.39 0.000 Observation Order
2
Analysis of Variance
Source DF SS MS F P 1
Regression 1 488.69 488.69 29.00 0.000
Residual Error 182 3066.44 16.85
Total 183 3555.13 0
-1
Regression Analysis: Bromate as BrO3 (ug/l) versus H(T-9)
Standardized Residual
-2
The regression equation is
Bromate as BrO3 (ug/l) = 13.6 - 0.531 H(T-9)
-3
184 cases used, 1096 cases contain missing values 01/01/2006 01/01/2007 01/01/2008 01/01/2009
Date
2
S = 4.16036 R-Sq = 11.4% R-Sq(adj) = 10.9%
1
Analysis of Variance
0
Source DF SS MS F P
Regression 1 404.96 404.96 23.40 0.000
Residual Error 182 3150.17 17.31 -1
Total 183 3555.13
Standardized Residual
-2
-3
Percent
R-Sq 21.7% 10 -2
R-Sq(adj) 21.2% 1 -3
Standardized Residual
15 0.1
-4 -2 0 2 4 0.9 1.0 1.1 1.2
Standardized Residual Fitted Value
Frequency
10
0 -3
0
Standardized Residual
0 1 2 3 4 5 6 7 8 9 -3.00 -2.25 -1.50 -0.75 0.00 0.75 1 0 0 0 0 0 0 0 0 0 0 0 0
10 20 30 40 50 60 70 80 90 100 110 120
H(T-5) Standardized Residual
Observation Order
Standardized Residual
-3
Analysis of Variance
-4
Source DF SS MS F P 01/01/2006 01/01/2007 01/01/2008 01/01/2009
Regression 1 1.4252 1.4252 35.63 0.000 Date
Residual Error 182 7.2803 0.0400
Total 183 8.7055
Residuals Versus 6600 - Lee Chalk
(response is Log(BrO3))
-1
-2
Standardized Residual
-3
-4
0 20 40 60 80 100 120 140
6600 - Lee Chalk
Regression Analysis: Bromate as BrO3 versus H(T-5), 6600 - Lee Chalk
Amwell Marsh Bromate V Hatfield Abstraction (T-5) & SMD-Lee
The regression equation is
Bromate as BrO3 (ug/l) = 16.5 - 0.567 H(T-5) - 0.0463 6600 - Lee Chalk Normal Probability Plot Versus Fits
99.9 3.0
99
184 cases used, 1096 cases contain missing values 1.5
90
50 0.0
Predictor Coef SE Coef T P
Percent
Constant 16.4972 0.5259 31.37 0.000 10 -1.5
H(T-5) -0.56654 0.09821 -5.77 0.000
6600 - Lee Chalk -0.046287 0.006332 -7.31 0.000 1
Standardized Residual
0.1 -3.0
-4 -2 0 2 4 5.0 7.5 10.0 12.5 15.0
S = 3.44667 R-Sq = 39.5% R-Sq(adj) = 38.9% Standardized Residual Fitted Value
Frequency
-1.5
6
Source DF Seq SS 0 -3.0
Standardized Residual
-1
Standardized Residual
-2
-3
01/01/2006 01/01/2007 01/01/2008 01/01/2009
Date
Hatfield bromate versus Hatfield abstraction S = 42.3414 R-Sq = 22.7% R-Sq(adj) = 22.5%
Regression Analysis: Bromate as BrO3 (ug/l) versus H(T)
Analysis of Variance
The regression equation is
Bromate as BrO3 (ug/l) = 347 - 11.4 H(T) Source DF SS MS F P
Regression 1 264752 264752 147.68 0.000
Residual Error 504 903568 1793
509 cases used, 771 cases contain missing values Total 505 1168321
Analysis of Variance
Predictor Coef SE Coef T P
Source DF SS MS F P Constant 317.162 3.893 81.47 0.000
Regression 1 320518 320518 190.28 0.000 H(T-4) -6.6668 0.7293 -9.14 0.000
Residual Error 506 852337 1684
Total 507 1172855
S = 44.6489 R-Sq = 14.3% R-Sq(adj) = 14.1%
100
400
Residual
300 0
100 -200
0 1 2 3 4 5 6 7 8 9 0 20 40 60 80 100 120 140
H(T) 6140 - Chilterns - East - Colne
Hatfield Bromate V Hatfield Abstraction Regression Analysis: Bromate as BrO3 versus H(T), 6140 - Chilterns
Normal Probability Plot Versus Fits
The regression equation is
99.9 400
99 Bromate as BrO3 (ug/l) = 338 - 12.5 H(T) + 0.245 6140 - Chilterns - East - Colne
90
200
50
509 cases used, 771 cases contain missing values
Percent
Residual
10
0
1
0.1 -200 Predictor Coef SE Coef T P
-200 0 200 400 250 275 300 325 350
Residual Fitted Value
Constant 338.259 4.644 72.85 0.000
H(T) -12.5002 0.7939 -15.75 0.000
Histogram Versus Order 6140 - Chilterns - East - Colne 0.24523 0.04290 5.72 0.000
150 400
Residual
50
Frequency
Analysis of Variance
0 -200
-160 -80 0 80 160 240 320 1 00 00 00 00 00 00 00 00 00 00 00 00
Residual
1 2 3 4 5 6 7 8 9 10 11 12 Source DF SS MS F P
Observation Order Regression 2 392546 196273 126.15 0.000
Residual Error 506 787276 1556
Total 508 1179822
Residuals Versus Date
(response is Bromate as BrO3 (ug/l))
100
Residuals from Bromate as BrO3 (ug/l) vs 6140 - Chilterns - East - Colne
Residual
0
-100
-200
01/01/2005 01/01/2006 01/01/2007 01/01/2008 01/01/2009
Date
Hatfield Bromate V Hatfield Abstraction & SMD-Colne
Chadwell Spring versus Hatfield abstraction
Normal Probability Plot Versus Fits
99.9
99 300 Regression Analysis: Bromate as BrO3 (ug/l) versus H(T)
90 200
Percent
0 Bromate as BrO3 (ug/l) = 1.58 - 0.0996 H(T)
Residual
10
-100
1
0.1
-100 0 100 200 300 250 275 300 325 350 141 cases used, 1139 cases contain missing values
Residual Fitted Value
Residual
Frequency
40 S = 0.871765 R-Sq = 8.0% R-Sq(adj) = 7.3%
-100
0
-160 -80 0 80 160 240 320 1 00 00 00 00 00 00 00 00 00 00 00 00
1 2 3 4 5 6 7 8 9 10 11 12
Residual
Observation Order
Analysis of Variance
Source DF SS MS F P
Residuals Versus Date Regression 1 9.1432 9.1432 12.03 0.001
(response is Bromate as BrO3 (ug/l)) Residual Error 139 105.6365 0.7600
400 Total 140 114.7797
200 Bromate
as BrO3
Obs H(T) (ug/l) Fit SE Fit Residual St Resid
100 292 1.97 3.3000 1.3881 0.0960 1.9119 2.21R
607 4.35 3.2000 1.1505 0.0737 2.0495 2.36R
Residual
0
614 3.04 3.7000 1.2817 0.0798 2.4183 2.79R
621 3.96 3.2000 1.1901 0.0736 2.0099 2.31R
627 3.96 3.0000 1.1901 0.0736 1.8099 2.08R
-100 943 6.00 3.2000 0.9870 0.0910 2.2130 2.55R
950 5.99 3.5000 0.9872 0.0910 2.5128 2.90R
965 5.99 3.3000 0.9871 0.0910 2.3129 2.67R
-200
1008 5.99 3.5000 0.9875 0.0909 2.5125 2.90R
01/01/2005 01/01/2006 01/01/2007 01/01/2008 01/01/2009
1014 3.99 3.4000 1.1868 0.0735 2.2132 2.55R
Date
R denotes an observation with a large standardized residual.
Residuals Versus 6140 - Chilterns - East - Colne
(response is Bromate as BrO3 (ug/l))
Regression Analysis: Bromate as BrO3 (ug/l) versus H(T-1)
400
The regression equation is
300 Bromate as BrO3 (ug/l) = 1.51 - 0.0861 H(T-1)
200
141 cases used, 1139 cases contain missing values
100
Predictor Coef SE Coef T P
Residual
Constant 1.5104 0.1328 11.38 0.000
0 H(T-1) -0.08607 0.02814 -3.06 0.003
-100
S = 0.879593 R-Sq = 6.3% R-Sq(adj) = 5.6%
-200
0 20 40 60 80 100 120 140 Analysis of Variance
6140 - Chilterns - East - Colne
Source DF SS MS F P H(T-4) -0.10451 0.02732 -3.83 0.000
Regression 1 7.2375 7.2375 9.35 0.003
Residual Error 139 107.5421 0.7737
Total 140 114.7797 S = 0.864349 R-Sq = 9.5% R-Sq(adj) = 8.9%
Source DF SS MS F P
The regression equation is
Regression 1 10.933 10.933 14.63 0.000
Bromate as BrO3 (ug/l) = 1.45 - 0.0745 H(T-2)
Residual Error 139 103.847 0.747
Total 140 114.780
141 cases used, 1139 cases contain missing values
S = 0.885578 R-Sq = 5.0% R-Sq(adj) = 4.3% 140 cases used, 1140 cases contain missing values
Source DF SS MS F P
The regression equation is
Regression 1 13.063 13.063 17.75 0.000
Bromate as BrO3 (ug/l) = 1.53 - 0.0891 H(T-3)
Residual Error 138 101.587 0.736
Total 139 114.650
141 cases used, 1139 cases contain missing values
S = 0.877371 R-Sq = 6.8% R-Sq(adj) = 6.1% 140 cases used, 1140 cases contain missing values
Source DF SS MS F P
The regression equation is
Regression 1 13.604 13.604 18.58 0.000
Bromate as BrO3 (ug/l) = 1.63 - 0.105 H(T-4)
Residual Error 138 101.046 0.732
Total 139 114.650
141 cases used, 1139 cases contain missing values
Percent
S = 0.871130 R-Sq = 8.7% R-Sq(adj) = 8.0% 10
-1
1
Standardized Residual
0.1
Analysis of Variance -4 -2 0 2 4 0.6 0.9 1.2 1.5 1.8
Standardized Residual Fitted Value
Source DF SS MS F P
Regression 1 9.9265 9.9265 13.08 0.000
Residual Error 138 104.7238 0.7589 Histogram Versus Order
Total 139 114.6503 24 3
2
18
Regression Analysis: Bromate as BrO3 (ug/l) versus H(T-8) 1
12
0
The regression equation is
Frequency
6
Bromate as BrO3 (ug/l) = 1.65 - 0.119 H(T-8) -1
Standardized Residual
0
-1.50 -0.75 0.00 0.75 1.50 2.25 3.00 1 00 00 00 00 00 00 00 00 00 00 00 00
140 cases used, 1140 cases contain missing values Standardized Residual
1 2 3 4 5 6 7 8 9 10 11 12
Observation Order
Predictor Coef SE Coef T P
Constant 1.6474 0.1373 12.00 0.000
H(T-8) -0.11877 0.02932 -4.05 0.000
Residuals Versus Date
S = 0.861689 R-Sq = 10.6% R-Sq(adj) = 10.0% (response is Bromate as BrO3 (ug/l))
3
Analysis of Variance
Source DF SS MS F P
Regression 1 12.184 12.184 16.41 0.000 2
Residual Error 138 102.466 0.743
Total 139 114.650
0
Standardized Residual
-1
-2
01/01/2006 01/01/2007 01/01/2008 01/01/2009
Date
The regression equation is
Residuals Versus 6600 - Lee Chalk log10(Bromate) = 0.135 - 0.0438 H(T-6)
(response is Bromate as BrO3 (ug/l))
140 cases used, 1140 cases contain missing values
3
Analysis of Variance
0 Source DF SS MS F P
Regression 1 1.8348 1.8348 18.72 0.000
Residual Error 138 13.5246 0.0980
Standardized Residual
Total 139 15.3593
-1
50
0
Percent
-1
10
Chadwell Spring Bromate V Hatfield Abstraction (T-6) -2
1
Bromate as BrO3 (ug/l) = 1.684 - 0.1194 H(T-6)
Standardized Residual
0.1
-4 -2 0 2 4 -0.3 -0.2 -0.1 0.0 0.1
4 Regression Standardized Residual Fitted Value
95% CI
95% PI Histogram Versus Order
3 S 0.855697 24 2
R-Sq 11.9%
18 1
R-Sq(adj) 11.2%
2 0
12
-1
Frequency
6
-2
1
Standardized Residual
0
-2.25 -1.50 -0.75 0.00 0.75 1.50 2.25 1 00 00 00 00 00 00 00 00 00 00 00 00
1 2 3 4 5 6 7 8 9 10 11 12
Standardized Residual
Observation Order
0
0 1 2 3 4 5 6 7 8 9
H(T-6)
Analysis of Variance
0
Source DF SS MS F P
Regression 2 2.1510 1.0755 11.16 0.000
Residual Error 137 13.2083 0.0964
-1 Total 139 15.3593
Standardized Residual
-2 log10(Chadwell Spring Bromate) V Hatfield Abstraction (T-6) & SMD-Lee
Normal Probability Plot Versus Fits
99.9
-3 2
99
01/01/2006 01/01/2007 01/01/2008 01/01/2009
90
Date 0
50
Percent
10
5
-2
0
Standardized Residual
Standardized Residual
-2
-3
0 20 40 60 80 100 120 140
6600 - Lee Chalk
-1
Standardized Residual
-2
-3
01/01/2006 01/01/2007 01/01/2008 01/01/2009
Date
284
Appendix D
D.1 Introduction
A series of single borehole dilution tests (SBDT) within existing boreholes in the Hertfordshire Chalk
were undertaken during 2008. These single borehole dilution tests were undertaken in conjunction with
a programme of point-to-point Natural Gradient Tracer Testing using Bacteriophage which are described
by Cook (2010).
D.2 Objectives
The objectives were:
• To determine the hydraulically active horizons within the selected boreholes in order to guide
injection strategies for the natural gradient point-to-point tracer testing;
• To use uniform injection SBDTs to obtain a direct measurement of horizontal specific discharge
(Darcy velocity);
• To use point injection SBDTs to determine vertical flow rates within the boreholes;
• Geophysical logging (Temperature & Conductivity, Gamma & Caliper, Formation Resistivity and
Impellor Flowmeter)
• Down-hole CCTV
Four boreholes were identified as suitable based on criteria outlined by Maurice (2009):
The location at Hatfield Business Park was not tested due to difficulties in obtaining permission.
D.3 Methodology
The field methodology followed Ward et al. (1998) and Maurice (2009). To summarise, a concentrated
solution of table salt (sodium chloride) was injected as a tracer into the water column via a weighted
hosepipe. For uniform injections, tracer solution was injected throughout the full length of water column;
for point injections, tracer solution was added at a specific interval. A Solinst Levelogger LTC probe was
used to measure the electrical conductance profile before and after injection to monitor the dilution of
the salt within the water column to background concentrations over time. The probe measured depth
below the water table using a pressure transducer. Corrections were made for changes in atmospheric
pressure, and temperature adjustment periods.
A proforma and groundwater risk assessment was submitted to the Environment Agency for ap-
proval in November 2007.
Figure D.1: Borehole construction details for the boreholes used for the single borehole dilution testing.
D.4.1 Nashe’s Farm BH
The borehole construction and water level details are shown schematically in Figure D.1.
• The calliper trace indicates that the open hole section of the borehole has a variable diameter, with
some fractures/fissures widening to 300 mm. In particular there are increases in borehole diameter
between 20.5 and 22.5 m below datum (bD) and between 24.5 and 25.5 m bD.
• The impellor flow logs detected some upward flow within the borehole between 23.3 m bD and
24.8 m bD.
• The impeller flowmeter logs detected a slight downward flow within the borehole from 17.0mbD,
in the slotted section of casing.
The EC profile immediately after injection shows elevated concentrations between ∼21.5 and
∼25.0 m bD in addition to the high concentrations seen at the bottom of the borehole in the back-
ground profile and subsequent profiles. Subsequent profiles show that the tracer front moves up the
borehole (vertical flow) until it reaches ∼17.0 m bD at 26 minutes. The summed tracer concentrations
Figure D.4: Electrical Conductance profile for uniform injection (column dilution) SBDT and point injection SBDT at Harefield House BH. Two point injections were carried
out: one with tracer injected at 22.9 m bD and a separate one with tracer injected at 19.5 m bD. Vertical flow rate is based on the vertical position of the point injection
concentration peak.
indicate that the mass appears to increase from the initial profile until the profile at 26 minutes. There
do not appear to be any consistent ‘nick points’ which would indicate outflow horizons over the section
of vertical flow.
The EC profile immediately after injection shows elevated concentrations between ∼81.5 and
∼22.5 m bD, with a relatively sharp peak at ∼20.5 m bD. Subsequent profiles show that the tracer peak
moves up the borehole (vertical flow) as it dilutes, until it returns to background concentrations by 37
minutes. There do not appear to be any consistent ‘nick points’ which would indicate outflow horizons
over the section of vertical flow. Summed tracer concentrations for point injection 2indicate mass loss
over the first 16 minutes as the peak moves between ∼20.5 and ∼17.0 m bD. Summed concentrations
hover around background concentrations for the remainder of the dilution until 74 minutes.
D.5.1 Methodology
Methodology is outlined in Figure D.5
D.5.2 Results
Ct −Cb
Plots of ln C0 −Cb
for Nashe’s Farm, Harefield House, and Comet Way are included.
Figure D.5: Methodology for determination of specific discharge (darcy velocity) from the results of the
Single Borehole Dilution Tests.
Harefield House 16.0 mbD Harefield House 16.50 mbD Harefield House 17.00 mbD
time (minutes) time (minutes) time (minutes)
0 50 100 150 200 250 300 350 0 50 100 150 200 250 300 350 0.0 50.0 100.0 150.0 200.0 250.0 300.0 350.0
0.0 0.0 0.0
y = -0.0122x - 1.0401 y = -0.0135x - 0.4139 y = -0.005x - 0.3944
-0.5 R2 = 0.6068 -0.5 2 -0.5 2
R = 0.7228 R = 0.3279
-1.0 -1.0 -1.0
-1.5 -1.5 -1.5
ln(Ct - Cb/Co - Cb)
-6.0
-6.0 -6.0 -8.0
-10.0
-8.0 -8.0
-12.0
-10.0 -10.0
-14.0
-12.0 -12.0 -16.0
Comet Way 18.40 mbD Comet Way 18.60 mbD Comet Way 18.80 mbD
time (minutes) time (minutes) time (minutes)
0 20 40 60 80 100 120 140 0 20 40 60 80 100 120 140 0 20 40 60 80 100 120 140
0.0 0.0 0.0
y = -0.1071x - 0.0921 y = -0.1106x - 0.0669 y = -0.1014x - 0.184
2 2
-2.0 R = 0.9913 -2.0 R = 0.9937 -2.0 R2 = 0.9772
-4.0 -4.0 -4.0
ln(Ct - Cb/Co - Cb)
-6.0
-4.0 -2.0
-3.0
-2.0 -2.0
Dear Ciara/Simon
Please find below a summary of the downhole inspection undertaken in Borehole No. 5 at Comet Way,
Hatfield, conducted on Tuesday 29th January.
K:\CMF Thesis\Appendices\Jess_Memos\CometWay_Memo.doc
1
Figure 2: Borehole Chamber
Geophysical Logging
Figure 3 displays the fully processed composite log of the geophysical tools ran during the survey.
The fluid temperature and conductivity sonde was not used during the inspection as there was only
3.8m of water in the borehole. The tool itself is over two and a half metres long. The borehole is
cased to total depth, therefore the formation resistivity tool could not be run. During logging, the
legs on the caliper sonde did not open. This is likely to be due to the poor verticality seen in the
borehole.
The filtered gamma log displays a count rate of 80 counts per second (CPS) until 2.8mbD,
suggesting a clay rich soil or made ground overlying the Chalk. Beneath here, the rate decreases to
19CPS, suggesting a lithological boundary with the Chalk. At 12.7mbD, the count rate is seen to
increase again to a maximum of 50CPS before falling when the water level is encountered, which is
to be expected as the gamma reading is attenuated when measured in water. The final two metres of
the profile displays a constant value of 11CPS.
The impeller flowmeter logs detected a slight downward flow within the borehole from 17.0mbD,
in the slotted section of casing. Flow horizons, if present, are indicated by changes in the impeller
rotation rate and time per revolution that are not associated with changes in the cable speed.
K:\CMF Thesis\Appendices\Jess_Memos\CometWay_Memo.doc
2
Well Name: BH 5
File Name: S:\WATERR~1\TVW\CENTRAL\BROMAT~1\DATA\DOWNHO~1\COMETW~1\BH5COMP.HDR
Location: Comet Way Datum: Ground Level
Metres Filtered Gamma Down Cable speed Down rotation rate Down time per rev Up Cable speed Up rotation rate Up time per rev
0 (CPS) 200 0 (m/min) 1000 0 (RPM) 80 0 (ms) 4000 0 (m/min) 1000 0 (RPM) 100 0 (ms) 3000
0
-5
-10
-15
-20
K:\CMF Thesis\Appendices\Jess_Memos\CometWay_Memo.doc
3
CCTV Inspection
A summary of the features seen during the CCTV inspection is shown in Table 1.
Depth Features
(mbD)
0 Top of Casing
1.2 Plain steel casing, in good condition
2.4 Casing joint, poor contact between sections, overhang present
8.1 Casing joint, poor contact between sections, overhang present
14.0 Casing joint, poor contact between sections, overhang present
CCTV camera knocks against the sides of the borehole. Beneath this
depth, borehole is off vertical
15.8 Rest Water Level, surface is dirty, water underneath is clear
16.9 Casing is slotted. Slots are clean, horizontal and arranged in three
columns. Off white Chalk is visible through the slots
19.4 Beer can is present in the borehole.
Estimated base of borehole is at 19.6mbD. Base is even and covered with
sediment
Regards
Jessica Randle
Water Resources
(01923) 814249/ #4249
K:\CMF Thesis\Appendices\Jess_Memos\CometWay_Memo.doc
4
Memorandum Date: 12/02/08
Dear Ciara/Simon
Please find below a summary of the downhole inspection undertaken at Nashes Farm, conducted on
Monday 4th February.
K:\CMF Thesis\Appendices\Jess_Memos\Nashes_Memo.doc
1
Figure 2: Borehole Chamber
Geophysical Logging
Figure 3 shows the geophysical logs produced by the fluid temperature and conductivity sondes run
in the borehole. During logging, the depth scale did not record properly, although all other
parameters are ok. Note the top of the log is at 19.6mbD and the base of the log is at 32mbD, giving
a 12.4m section. Figure 4 displays the fully processed composite log of the remaining geophysical
tools ran during the survey.
The fluid temperature profile displays 11.2C at the rest water level (RWL), which is encountered at
19.6mbD and remains constant until the base of the log. Sudden changes in the geothermal gradient
may be due to possible flow horizons. The differential temperature log indicates that no discernable
fluctuations are evident within the borehole.
The fluid conductivity profile gradually decreases from 704S/cm at the RWL to 644S/cm at the
base of the log. Again, changes in the profile may infer flow horizons. The differential conductivity
log indicates no significant fluctuations relating to flow.
After initially high values reflecting the pumping house and surrounding made ground, the filtered
gamma log displays a count rate of 2.7 counts per second (CPS) until 8.7mbD. Beneath here, the
rate increases slightly to 3.6CPS. A decrease to 2.4CPS is observed from the rest water level
(RWL) at 19.6mbD, which is to be expected as the gamma reading is attenuated when measured in
water. This value is maintained until the base of the borehole. No obvious lithological boundaries
are identifiable from the gamma profile.
K:\CMF Thesis\Appendices\Jess_Memos\Nashes_Memo.doc
2
The caliper trace confirms the internal diameter of the plain casing to be 160mm. The open hole
section of the borehole appears well fractured and displays a variable diameter, with some fissures
widening to 300mm.
The formation resistivity varies between a minimum of 38.5m at 21.6mbD and 29.3m at
23.5mbD. With only 10m of saturated open hole section, it is difficult to stratigraphically place
where the borehole lies in terms of Chalk formations. There are no obvious marker horizons present
and the resistivity trace is strongly linked to fluctuations with the caliper log. The shallow depth of
the borehole, the well fissured nature of the open hole section and the presence of large sections of
flint would suggest that the borehole comprises Upper Chalk. The sudden increase in resistivity at
the bottom of the log is due to a test measurement carried out during logging to ensure the sonde is
calibrated correctly.
The impeller flowmeter logs detected some upwards flow within the borehole between 23.3mbD
and 24.8mbD. Flow horizons, if present, are indicated by changes in the impeller rotation rate and
time per revolution that are not associated with changes in the cable speed.
Well Name:
File Name: S:\WATERR~1\TVW\CENTRAL\BROMAT~1\DA
Location: Nashes Farm Datum: Pump house Floor
Fluid Temp Diff Fluid Temp Fld Cond @ 25°C Diff Fld Cond
8 (°C) 16 -1 (°C/m) 1 400 (µS/cm) 800 -60 (µS/cm/m) 60
K:\CMF Thesis\Appendices\Jess_Memos\Nashes_Memo.doc
3
Well Name:
File Name: S:\WATERR~1\TVW\CENTRAL\BROMAT~1\DATA\DOWNHO~1\NASHES~1\NASHCOMP.HDR
Location: Nashes Farm Datum: Pump House Floor
Metres Filtered Gamma BH Diameter Form Resistivity Down Cable speed Down rotation rate Up Cable speed Up rotation rate
0 (CPS) 40 0 (mm) 400 0 (Ohm-m) 100 0 (CPM) 1000 0 (RPM) 60 0 (CPM) 1000 0 (RPM) 60
0
-5
-10
-15
-20
-25
-30
Figure 4: Nashes Farm Borehole Composite Geophysical Log
CCTV Inspection
A summary of the features seen during the CCTV inspection is shown in Table 1.
Depth Features
(mbD)
0 Pump House floor
1.2 Plain steel casing, slightly rusted, in reasonable condition
5.9 Base of casing
Chalk is stained brown, smooth and stable
K:\CMF Thesis\Appendices\Jess_Memos\Nashes_Memo.doc
4
Depth Features
(mbD)
7.4 Borehole circumference appears irregular
8.0 Chalk opens out to one side, flints present
9.3 Borehole circumference appears irregular
11.6-12.3 Chalk surface becomes more rough
13.8 Chalk opens out to one side. Beneath here the Chalk is whiter
16.5 Chalk opens out to one side
17.3 CCTV camera gets stuck on a large piece of flint protruding from the side
of the borehole
18.3 Borehole circumference appears irregular
19.5 Rest water level, water is reasonably clear
19.6 Chalk appears smooth
20.4 Borehole opens out and the circumference appears irregular
22.5 CCTV camera knocks against a ledge on the side of the borehole.
Chalk beneath here is smooth
23.1-24.4 Large vertical groove appears to one side of the borehole
24.4 Borehole opens out, circumference becomes irregular
30.7 CCTV Camera gets stuck on the sides of the borehole, survey terminated
*Please note the borehole was plumbed to a depth of 32.70m
Regards
Jessica Randle
Water Resources
(01923) 814249/ #4249
K:\CMF Thesis\Appendices\Jess_Memos\Nashes_Memo.doc
5
285
Appendix E
Figure 1: Comparison of simulated mobile zone concentrations (no matrix diffusion) using GSIM and MAP.
Concentrations in the immobile zone were simulated using a ‘diffusion cells’ approach (see note
below). Since MAP is not currently coded to give immobile zone concentrations, the simulated
concentrations could not be validated against MAP.
Figure 2: Comparison of simulated mobile zone concentrations with matrix diffusion using GSIM and MAP.
The single streamtube modelled with DP1D is an average of the 9 streamtubes leading to the output
‘Sink 2’. This was chosen so that an approximate comparison could be made between this
GSIM/DP1D single streamtube and MAP output for Sink 2.
Despite using the same parameters in both simulations, simulated mobile and immobile zone
concentrations are significantly different (Figure 3; Figure 4). . (Mobile zone concentrations remain
very similar in GSIM and MAP outputs). In particular, the relative difference between mobile and
immobile concentrations is larger in DP1D than in GSIM, indicating that more diffusion into the
matrix is occurring in GSIM than in DP1D (Figure 4).
Figure 3: Comparison of mobile zone concentrations with matrix diffusion using GSIM and MAP & DP1D.
Figure 4: Comparison of mobile and immobile zone concentrations using GSIM, MAP & DP1D. Solid red line is
DP1D mobile concentration, dashed red line is DP1D immobile concentration. Other lines represent GSIM immobile zone
concentration at distances into the matrix block, until half block thickess.
Possible explanations that were considered are summarised below:
2.1.4 Conclusions
Simulations with GSIM and MAP, using identical/best‐estimate parameters, produce identical/similar
output concentrations for mobile concentrations with and without inclusion of matrix diffusion.
However, it is not possible to compare simulated immobile zone concentrations as neither model is
coded to produce this as an output.
An approximation for immobile zone concentrations can be made by setting up a series of coupled
‘diffusion cells’ which are linked to the output mobile zone concentration. However, there is no
MAP output to validate/verify this approach. When a single pathway is simulated with DP1D and
GSIM, mobile and immobile zone concentrations are substantially different.
NOTE: ‘Diffusion cells’ approach to estimate matrix concentrations in GSIM
This approach is represented diagrammatically in Figure 5.
A series of GSIM ‘cell pathways’ (mixing cells) are generated to represent half a matrix block. The
first cell contains water only, and its concentration is fixed to the output concentration from the pipe
pathway (streamtube) mobile zone at that point. The rest of the cells contain saturated porous
matrix material and represent thin slices of matrix moving from the fracture into the matrix block.
Each matrix cell is coupled to the adjacent cells by a ‘diffusive mass transfer link’. This link
represents diffusive exchange of mass between immobile porewater in adjacent matrix cells, and is
described using a constant mass transfer rate. The linked cells therefore form a finite difference
network. Concentrations are calculated at each time step.
Figure 5: The ‘diffusion cells’ approach to estimate immobile concentrations in GSIM.
Diffusion Cell Set‐up
Describe set‐up
X1
Fracture Y1 0
Matrix Y2 0.001
Matrix Y3 0.003
Matrix Y4 0.005
Matrix Y5 0.01
Matrix Y6 0.02
Matrix Y7 0.03
Matrix Y8 0.05
Matrix Y9 0.1
The numbers in the cells specify the Y‐coordinate for the start of the corresponding cell element. In the last row
the number specifies the Y‐coordinate for the end of the last Cell.
Fracture cells contain water only
Matrix cells contain chalk with porosity and water within the pores.
The fracture cell has a ‘specified concentration’ boundary condition which is specified as the
concentration output of the pipe element.
Cells are linked by a diffusive mass transfer links.
The diffusive flux f from pathway i to pathway j is computed as follows:
f ij = D(ci – cj)
where D = diffusive conductance for the species in the mass flux link [L3/T], ci and cj are dissolved
concentration of the species in the medium (water) within cell i and cell j respectively.
Diffusive conductance terms are computed as follows:
D = A
Li + Lj
d*tpi*npi d*tpj*npj
where Li and Lj are the diffusive lengths for the diffusive mass flux link in cell i and cell j (the default
is distance from the centre of the cell to the edge or interface of the cell)
d is the diffusivity for species in water
tpi is the tortuosity for the porous medium in cell i
npi is the porosity for the porous medium in cell i
GoldSim calculates the effective diffusion coefficient for the pipe pathway as Dim = d*t*n.
I have set d, t and n to give a Dim of 1.74 x 10‐10 m2/s, which is the value in DP1D, therefore n = 0.5,
t = 0.7 and d = 0.50 x 10‐9 m2/s.
286
Appendix F
Join to Arkley Hole (Karst) 2.60E+00 1.65E+00 4.38E+05 1.90E+01 1209.125 0.001
Harefield House to Arkley Hole (Karst) 3.04E-03 2.30E+01 4.38E+05 1.90E+01 1209.125 0.01
Harefield House to Lynchmill (Karst) 1.31E-02 5.29E+01 9.04E+04 1.90E+01 249.9371 0.01
Comet Way to Arkley Hole (Karst) 8.23E-01 9.02E+00 4.38E+05 1.90E+01 1209.125 0.01
Comet Way to Lynchmill (Karst) 1.61E+00 9.02E+00 9.04E+04 1.90E+01 249.9371 0.01