opus.bath.ac.ukopus.bath.ac.uk/53573/1/Collender_disease_risks_inla… · Web viewopus.bath.ac.uk

111
Estimating the microbiological risks associated with inland flood events: Bridging theory and models of pathogen transport Philip A. Collender 1 , Olivia C. Cooke 2 , Lee D. Bryant 2 , Thomas R. Kjeldsen 2 and Justin V. Remais 1 1 Environmental Health Sciences, School of Public Health, University of California, Berkeley, Berkeley, CA 94720 2 Department of Architecture and Civil Engineering, University of Bath, Bath, UK BA2 7AY 1 1 2 3 4 5 6 7 8 9 10 11 12

Transcript of opus.bath.ac.ukopus.bath.ac.uk/53573/1/Collender_disease_risks_inla… · Web viewopus.bath.ac.uk

Page 1: opus.bath.ac.ukopus.bath.ac.uk/53573/1/Collender_disease_risks_inla… · Web viewopus.bath.ac.uk

Estimating the microbiological risks associated with inland flood events: Bridging theory and models of pathogen transport

Philip A. Collender1, Olivia C. Cooke2, Lee D. Bryant2, Thomas R. Kjeldsen2 and Justin V. Remais1

1 Environmental Health Sciences, School of Public Health, University of California, Berkeley, Berkeley, CA 94720

2 Department of Architecture and Civil Engineering, University of Bath, Bath, UK BA2 7AY

1

123

4

5

6789

10

Page 2: opus.bath.ac.ukopus.bath.ac.uk/53573/1/Collender_disease_risks_inla… · Web viewopus.bath.ac.uk

Abstract:

Flooding is known to facilitate infectious disease transmission, yet quantitative research on microbial

risks driven by floods has been limited. Pathogen fate and transport models provide a framework for

examining impacts of landscape characteristics and hydrology on infectious disease, but have not been

widely developed for flood conditions. We critically examine capacities of existing hydrological models

to represent the unusual flow paths, non-uniform flow depths, and unsteady flow velocities accompanying

flooding. We investigate theoretical linkages between hydrodynamic processes and spatio-temporally

variable suspension and deposition of pathogens from soils and sediments, pathogen dispersion in flow,

and concentrations of constituents that influence pathogen transport and persistence. Identifying gaps in

knowledge and modeling practice, we propose a research agenda to strengthen microbial fate and

transport modeling applied to inland floods: 1) development of models incorporating pathogen discharges

from flooded latrines, effects of transported constituents on pathogen persistence, and supply-limited

pathogen transport; 2) studies assessing parameter identifiability and comparative performance of models

with varying degrees of process representation in various settings; 3) development of remotely sensed

datasets for models of vulnerable, data-poor regions; and 4) collaboration between modelers and field-

based researchers to expand the collection of useful data in situ.

2

11

12

13

14

15

16

17

18

19

20

21

22

23

24

25

26

Page 3: opus.bath.ac.ukopus.bath.ac.uk/53573/1/Collender_disease_risks_inla… · Web viewopus.bath.ac.uk

1. Floods and infectious disease transmission

Floods are a major, recurring source of harm to global economies and public health. In the past decade,

floods have caused US$185 billion in economic losses and have been responsible for 65,000 deaths

globally, accounting for nearly half of the mortality associated with natural disasters (EM-DAT 2014).

Future changes to the global climate and increasing urbanization may exacerbate the impact of floods.

Projected increases in the frequency and intensity of heavy precipitation events in some areas, for

instance, may expose larger populations to more frequent and severe flooding in future decades (Jonkman

2005, United Nations 2012, IPCC 2013). Flood-related mortality is greatest in populations with poor

infrastructure and limited economic resources; thus, the impact of flooding in developing countries in

particular is set to remain a major source of morbidity, mortality and economic loss in the coming

decades (Ahern, et al. 2005). Improved flood risk management is therefore essential to support global

development, protect infrastructure, and improve public health.

When considering the public health impact of floods in a risk management framework, most analyses

consider only a subset of the direct pathways linking flood events to mortality and morbidity (e.g.,

injuries, drowning), neglecting more subtle and sometimes delayed impacts such as those resulting from

infectious diseases (Jonkman, et al. 2008). At the same time, it is well known that flood conditions—

defined in detail below—can exacerbate the spread of infectious diseases through various mechanisms

affecting the transport and persistence of pathogenic microorganisms. Microbial contamination of surface

water can result from flood conditions following heavy rainfall when high flow volumes and suspended

solids overwhelm water treatment systems, or when fecal material is flushed from sources on the land

surface due to runoff or inundation by an overflowing channel (Ahern, et al. 2005, Hunter 2003,

Alderman, et al. 2012). High flow velocities in flooded channels may lead to resuspension of pathogens

persisting in channel bed sediments, which may then be transported over long distances in rapidly flowing

surface water (Hunter 2003, McBride and Mittinty 2007). Furthermore, increased transport of suspended

3

27

28

29

30

31

32

33

34

35

36

37

38

39

40

41

42

43

44

45

46

47

48

49

50

51

Page 4: opus.bath.ac.ukopus.bath.ac.uk/53573/1/Collender_disease_risks_inla… · Web viewopus.bath.ac.uk

solids in flooded channels may affect microbial persistence, potentially allowing pathogens to remain

infectious as they reach locations far from their original sources (Walters, et al. 2014).

1.1 Models for characterizing flow and microbial dynamics during flooding

Integrated modeling platforms are available to characterize the timing and extent of flooding during

heavy rainfall events; these generally consist of rainfall-runoff models for estimating channel flows in

response to precipitation inputs, followed by hydrodynamic modeling to simulate the movement of water

through channels and floodplains. Rainfall-runoff models provide a quantitative representation of the

interaction between meteorology and channel flows, and of the partitioning of water between

environmental compartments (e.g., surface storage, subsurface storage, and channel storage), and may

take into account topology, soil type, vegetation and other land-surface characteristics of entire upstream

drainage areas. Rainfall-runoff models may also be used to examine the effects of changes in climate and

land management on the frequency and magnitude of flood events as reflected by their hydrograph

outputs (Singh and Woolhiser 2002, Miller, et al. 2014). Hydrodynamic models are used to route flood

waves (i.e., track water depth and velocity of flow through space and time) through channels and

floodplains based on principles of conservation of mass and, in many applications, momentum (Singh

1996){Singh, 1996 #80;Singh, 1996 #12}.

In most practical applications, a modeler will initially choose a model system deemed suitable for the task

at hand. Next, model parameters are calibrated by tuning the parameter values until as good a fit between

observed and simulated flows, water levels, and/or inundation extents can be obtained. Ideally, some

observed data are withheld from the calibration process and used for subsequent model validation to

ensure that the model performs as expected outside the range of the calibration data (Klemeš 1986). In

practice, rainfall-runoff and hydrodynamic models are affected by uncertainty, parameter non-

identifiability, and model equifinality (a phenomenon in which multiple structurally distinct models

perform similarly well after calibration to observed data) (Beven 2006). These issues are unavoidable in

4

52

53

54

55

56

57

58

59

60

61

62

63

64

65

66

67

68

69

70

71

72

73

74

75

76

77

Page 5: opus.bath.ac.ukopus.bath.ac.uk/53573/1/Collender_disease_risks_inla… · Web viewopus.bath.ac.uk

hydrological modeling, stemming from the complexity of the factors at play in determining hydrological

responses, the non-linear relationships between them, their spatiotemporal heterogeneity, mismatches in

scale between process descriptions and empirical equations derived from lab experiments, data available

from local or remote sensing, and model spatial discretization, and the limited availability, reliability, and

spatiotemporal coverage of observations used to parameterize or calibrate model process descriptions

(Beven 2006, Beven 2001). Despite these practical difficulties, hydrological and hydrodynamic models

provide a necessary framework for integrating knowledge on the many factors determining regional

hydrology and flood risk.

To estimate microbial hazards alongside physical hazards of floods, additional subcomponents can be

incorporated into hydrological and hydrodynamic models representing the mobilization, transport, and

fate of pathogens in runoff, channels and inundated areas. Microbial transport has been incorporated into

a number of rainfall-runoff and hydrodynamic models, as reviewed by Jamieson et al. (Jamieson, et al.

2005) and de Brauwere et al. (de Brauwere, et al. 2014). These models have primarily been developed for

and applied in agricultural settings in order to assess and attribute microbial pollution of receiving waters.

Issues of parameter non-identifiability and equifinality are exacerbated for fate and transport models, due

to the large number of additional parameters that may be incorporated into the model calculations, as well

as to the scarcity of spatially distributed measurements of microbial concentrations within a catchment

(Sommerfreund, et al. 2010). Issues of data scarcity are even more severe when analyzing flood events:

Gauge data is usually confined to channels (Neal, et al. 2009), leading to difficulty in obtaining data for

calibration of flow parameters in inundated areas, and microbiological measurements are especially

scarce during high flows (Ghimire and Deng 2013). This is probably why integration of microbial

transport modeling into analysis of flood events, especially the analysis of flows in inundated areas, is

notably sparse in the literature. In one example linking transport modeling to flood conditions, Wu et al.

(2009) used the Waterloo flood system distributed hydrological model (WATFLOOD) to predict the

temporal and spatial distribution of E. coli in the Blackstone River (USA) during a series of wet weather

5

78

79

80

81

82

83

84

85

86

87

88

89

90

91

92

93

94

95

96

97

98

99

100

101

102

103

Page 6: opus.bath.ac.ukopus.bath.ac.uk/53573/1/Collender_disease_risks_inla… · Web viewopus.bath.ac.uk

events in 2005-06. They attained correlation between modeled and observed in-stream bacterial

concentrations within an order of magnitude upstream of a wastewater treatment plant and combined

sewer overflow (CSO) discharge, though the model did not perform well downstream of the plant and

CSO. Ghimire and Deng (2013) augmented the Variable Residence Time (VART) model for bacterial

transport in stream flows to account for transient storage in permeable banks and sediment, and applied it

to several high flow events in the Motueka River, New Zealand. Estimates of in-stream E. coli

concentrations produced by their model matched observed concentrations reasonably well (r2 range: 0.36

- 0.9) over 12 storm events. Kazama et al. (2012, 2007) used a hydrodynamic model to perform a

quantitative microbial risk assessment for monsoon-driven flooding of the lower Mekong River,

Cambodia, considering transport in both channels and inundated areas. Notably, they found that areas

with heightened exposures to fecal coliforms simulated by their model also displayed elevated infant

mortality from diarrheal disease; these results suggest that pathogen transport modeling for flood events

can be used to understand waterborne disease risks.

6

104

105

106

107

108

109

110

111

112

113

114

115

116

117

Page 7: opus.bath.ac.ukopus.bath.ac.uk/53573/1/Collender_disease_risks_inla… · Web viewopus.bath.ac.uk

1.2 Unique features of microbial fate and transport during flood events

7

118

Page 8: opus.bath.ac.ukopus.bath.ac.uk/53573/1/Collender_disease_risks_inla… · Web viewopus.bath.ac.uk

Microbial fate and transport during flood events is affected by unusual flow paths taken through complex

topography by overbank flows, and by variations in flow depth and velocity over time and space that may

be more complex than those experienced under nominal flow conditions. Flow paths that extend across

inundated areas may mobilize pathogens from soils, feces, or flooded sanitation facilities (e.g., latrines)

on the land surface that might not be considered as sources of contamination under nominal runoff

conditions. Furthermore, areas affected by overbank flows or heavy rainfall may exhibit markedly altered

pathogen mobilization and transport, due to the activation of preferential flow pathways and decreased

retention of transported pathogens in the surface layers of saturated soils (Bradford, et al. 2013).

Meanwhile, flooding can give rise to variations in flow depth and velocity beyond what would be

experienced under normal conditions as a result of increased in-channel flow volume relative to the

wetted perimeter, variable depth, width, and direction of flow across inundated areas, and complex

transfers of momentum between areas of deeper and shallower flow (Costabile, et al. 2013, Ikeda and

McEwan 2009, Nittrouer, et al. 2011). In addition, backwater effects can occur along flooded channels or

across inundated areas, where alterations of upstream flow characteristics are induced by downstream

obstructions (e.g., rising waters become partially obstructed by a bridge), or transitions to sub-critical

flow regimes (Ikeda and McEwan 2009). In turn, flow depth and velocity directly impact the kinetic

energy available for pathogen mobilization (Wu, et al. 2009, Schulz, et al. 2009), mechanical dispersion

of pathogens along primary and secondary axes of flow (Elder 1959, Fischer 1975, Kashefipour and

Falconer 2002), and the mobilization of solids and organic matter from bed sediments and the land

surface. Increased transport of organic matter and suspended solids in flood waters may affect microbial

persistence and transport in complex ways: transported colloids may prolong microbial persistence by

partially blocking UV radiation and increasing nutrient concentrations in the water column (Walters, et al.

2014), dissolved organic carbon (DOC) and other nutrients may provide a favorable environment for

bacterial survival and regrowth – or decrease survival by promoting the growth and/or activity of

antagonistic microbes, and organic matter and surfactants may compete with microbes for binding sites in

soils and sediments, enhancing microbial mobility (Bradford, et al. 2013).

8

119

120

121

122

123

124

125

126

127

128

129

130

131

132

133

134

135

136

137

138

139

140

141

142

143

144

Page 9: opus.bath.ac.ukopus.bath.ac.uk/53573/1/Collender_disease_risks_inla… · Web viewopus.bath.ac.uk

In this manuscript, we review specific aspects of hydrodynamic models that impact the simulation of flow

paths (Section 2), flow depths and flow velocities (Section 3) under flood conditions. In Section 4, we

examine key aspects of pathogen transport models that affect model performance under flood conditions.

In tables presented throughout the review, and in our discussion section (Section 5), we critically examine

a representative set of rainfall-runoff and flow-routing models adapted for microbial transport estimation

with respect to their capacities to represent key processes affecting pathogen movement and persistence

under flood conditions. Additionally, we discuss key knowledge gaps relevant to fate and transport during

floods, tradeoffs between detailed process representation and model uncertainty, and the importance of

emerging data sources and enhanced collaboration between modelers and researchers in the field for

advancing understanding of microbial risks associated with flooding. We do not attempt a discussion of

all processes relevant to microbial fate and transport (for comprehensive reviews, see Jamieson et al

(2005), Bradford et al. (2013), and de Brauwere et al. (2014)), but limit our focus to aspects of fate and

transport that seem especially relevant to flood conditions.

2. Modeling flow paths during floods

During flood events, runoff and overbank flows pass through areas that are not wetted under normal

conditions as a result of surface water volumes exceeding the capacity of soils, channel banks, and

hydraulic structures to absorb or contain them. The capability of rainfall-runoff models to represent paths

taken by overland flow is determined at a fundamental level by their spatial resolution, while the

dimensionality of hydrodynamic models affects the routing of overbank flows.

The spatial resolution of rainfall-runoff models (Table 1) determines the specificity with which runoff and

transported pathogens may be routed from land surface and subsurface compartments into channels.

Spatially lumped models average landscape and climate characteristics and estimate a single set of

outputs over an entire drainage area. As a result, it is not straightforward to assess the reliability with

9

145

146

147

148

149

150

151

152

153

154

155

156

157

158

159

160

161

162

163

164

165

166

167

168

169

170

Page 10: opus.bath.ac.ukopus.bath.ac.uk/53573/1/Collender_disease_risks_inla… · Web viewopus.bath.ac.uk

which lumped models partition precipitation between runoff, interflow, and base flow components across

the region of interest, all of which have different implications for pathogen mobilization and transport.

Furthermore, no information on the actual path taken by runoff is generated within lumped models. As a

result, the path taken by pathogens in overland flow from sources to channels and spatial heterogeneity of

pathogen sources, runoff generation, and key processes of pathogen release, transport, and retention

across the basin cannot be explicitly represented. Rather, the total runoff generated is applied to the total

pathogens available for transport even if these parameters do not spatially coincide. Semi-distributed

models capture key aspects of a catchment’s hydrological response and land-surface processes, without

relying on a fully explicit spatial representation, and by lumping spatial elements into response units that

are thought to share common hydrological characteristics. Each response unit is treated as homogenous

with respect to processes generating runoff and transported pathogens (Bormann, et al. 2009). Semi-

distributed models can differentiate between the contributions of various land types within a catchment,

allowing for more realistic matching of runoff generation with pathogen mobilization. Fully distributed

models may either function similarly to semi-distributed models, applying empirical formulae to simulate

the amount of runoff generated from each spatial element (e.g., grid cells of a Digital Elevation Model),

or may represent overland flow between adjacent spatial elements using hydrodynamic equations. In the

latter case, the generation and routing of runoff and mobilized pathogens may be estimated across

overland flow paths.

While distributed hydrological models with flow routing between spatial elements may seem superior

because of the detail with which they represent the land surface and runoff generation processes, they are

prone to issues of parameter non-identifiability and equifinality. This is in large part due to the large

number of additional parameters which may be necessary to represent spatial heterogeneity, many of

which cannot be measured directly or at the scale relevant to the model, particularly the parameters of

subsurface processes (Beven 2001). Furthermore, it is widely acknowledged that the use of mathematical

models of open, complex and large-scale real-world environmental systems involves a range of necessary

10

171

172

173

174

175

176

177

178

179

180

181

182

183

184

185

186

187

188

189

190

191

192

193

194

195

196

Page 11: opus.bath.ac.ukopus.bath.ac.uk/53573/1/Collender_disease_risks_inla… · Web viewopus.bath.ac.uk

simplifications which introduce uncertainty into the modelling process (Beven 2010a). Thus, despite the

appearance of being more theoretically rigorous, complex distributed models do not necessarily guarantee

more reliable descriptions of hydrological systems, even when accompanied by an increase in the quality

of observational data available for calibration (Beven 2006, Beven 2001, Beven 2010a).

In the case of hydrodynamic models (Table 2), spatial dimensionality may constrain the representation of

flow paths in inundated areas. One-dimensional (1D) models perform best in areas with high volume and

narrow width of flow, as in most river reaches (Tayefi, et al. 2007). Two-dimensional (2D) models

usually provide an improved simulation of floodplain flows (Tayefi, et al. 2007) but require significantly

more computational resources (Pilotti, et al. 2014). Three-dimensional (3D) model formulations are

normally reserved for flow simulations in lakes, estuaries, and coastal waters. In many cases, modelers

have coupled 1D models for channel flow with 2D floodplain models, allowing for realistic routing of

floodplain flows while foregoing high computational costs associated with gridded 2D solutions for

channel flows (Bladé, et al. 2012). In 1D models, inundation is typically represented by a compound

channel in which water levels in overbank and channel areas are set as equivalent, with overbank flow

following the direction of channel flow but at a different velocity. Conversely, in fully 2D or coupled 1D

channel - 2D floodplain models, flow in inundated areas may be routed in any cardinal direction.

3. Modeling depth and velocity of flow during floods

Section 3 nomenclatureh Depth of flow

(m)t Time (continuous)

(units range from seconds to days)Q Discharge (rate of flow)

(L ∙ time-1)x Distance along primary axis of flow

(m)v Flow velocity

(m ∙ time-1)g Acceleration due to gravity

(m2 ∙ time-1)s0 Bed slope

(unitless)sf Friction slope (energy head loss)

(unitless)I Inflow

(L ∙ time-1)O Outflow

(L ∙ time-1)S Storage

(L)∆ t Time step (discrete)

(units range from seconds to days)

11

197

198

199

200

201

202

203

204

205

206

207

208

209

210

211

212

213

214

215

Page 12: opus.bath.ac.ukopus.bath.ac.uk/53573/1/Collender_disease_risks_inla… · Web viewopus.bath.ac.uk

α General term for user-defined or calibrated coefficient

β General term for user-defined or calibrated exponent

θ Muskingum weighting factor(unitless)

K s Ratio of storage to discharge in Muskingum equation(time)

During flood events, flow depth and velocity may vary substantially across time and space, which can

significantly influence microbial transport. Flow depth and velocity are key determinants of many

transport processes, including net mobilization of free and sediment-associated pathogens from surface

layers of soils and channel sediments; dispersion of entrained pathogens; and aspects of pathogen

persistence mediated by suspended solids and dissolved nutrients. Thus, various properties of rainfall-

runoff and flow-routing models that affect the estimation of flow parameters may impact simulations of

microbial transport. The temporal resolution of rainfall-runoff models (Table 1) used to generate

boundary conditions for flow models of flood-affected areas is important, as peak flow may be transient,

and models run at coarser time steps are known to provide less-accurate estimates of even daily discharge

during time periods encompassing storm events (Borah, et al. 2007). In open-channel hydrodynamic

models (Table 2), the dimensionality and physical fidelity of the equations for conservation of mass and

momentum that are used to simulate the movement of water across the model domain, as well as

techniques to simulate exchanges of flow between channels and inundated areas, and a modeler’s

implementation of spatial heterogeneities in roughness (a quantity representing losses of momentum to

friction or turbulence) may impact estimates of flow depth, velocity, and direction; we detail these in the

next section.

Table 1: Spatial and temporal resolution of rainfall-runoff microbiological transport

models

Hydrological model Minimum time step Spatial discretization References

HSPF 1 minute Semi-distributed (Bicknell, et al. 1996)

WATFLOOD 1 hour Semi-distributed (Wu, et al. 2009, Kouwen 2014, Dorner, et al. 2006)

IHACRES 1 day Lumped (Croke, et al. 2005, Ferguson, et al. 2007)

12

216

217

218

219

220

221

222

223

224

225

226

227

228

229

230

231

232

233

234

Page 13: opus.bath.ac.ukopus.bath.ac.uk/53573/1/Collender_disease_risks_inla… · Web viewopus.bath.ac.uk

SWAT 15 minutes Semi-distributed (Neitsch, et al. 2011, Jeong, et al. 2010)

KINEROS2 /STWIR 1 hour

Fully-distributed hydrodynamic

(kinematic wave routing)(2000, Guber, et al. 2014)

COLI 1 hour Lumped (Walker, et al. 1990)

WAMVIEW 1 day Fully-distributed conceptual (Soil & Water Engineering Technology Inc. , Tian, et al. 2002)

WEPP 1 secondFully-distributed hydrodynamic

(kinematic wave routing)

(Flanagan and Nearing 1995, Yeghiazarian, et al. 2006, Bhattarai, et al.

2011)

GIBSI 1 hour Semi-distributed (Simon, et al. 2013)

3.1 Hydrodynamic equations for conservation of mass and momentum

Many hydrodynamic models track the depth and velocity of flow through space and time using variations

of the de Saint Venant shallow-water equations for conservation of mass (eqn 1) and momentum (eqn 2),

presented here in 1D form (Singh 1996):

∂ h∂ t

+ ∂ Q∂ x

=0 (1 )

∂ v∂ t

+v ∂ v∂ x

+g ∂ h∂ x

+g(s0−sf )=0 (2 )

where h is water depth; t is time; Q is discharge; x is a positional coordinate in the direction of flow; v is

flow velocity; g is acceleration due to gravity; s0 is the channel bed slope; and sf is the friction slope.

Within the momentum equation (eqn 2), the terms account for local acceleration and unsteady flow ( ∂ v∂ t ),

convective acceleration (v ∂ v∂ x ), water pressure gradient (g ∂ h

∂ x ), gravitational acceleration along the

channel slope ( g s0 ) ,and momentum lost to friction (−g sf ), respectively. The complete form of the

momentum equation given above (eqn 2), also known as the dynamic wave equation, is capable of

modeling fully unsteady (time-varying), non-uniform (space-varying) flows, including the effect of

13

235

236

237

238

239

240

241

242

243

244

245

246

247

248

Page 14: opus.bath.ac.ukopus.bath.ac.uk/53573/1/Collender_disease_risks_inla… · Web viewopus.bath.ac.uk

downstream conditions on upstream depth and flow velocity (i.e., backwater effects). The dynamic wave

equation is considered the standard of physical fidelity for conservation of momentum, especially when

modeling overbank flows and sediment transport using high-resolution topographical data (Costabile, et

al. 2013).

Simplified approaches to conservation of momentum assume that one or more terms within the dynamic

wave equation may be considered negligible, resulting in model formulations that may assume space- or

time-invariant flows with varying capacity to represent backwater effects. Typically, the diffusive wave

(eqn 3) or kinematic wave (eqn 4) approximations may be used in place of the dynamic wave equation

(Miller 1984):

g ∂ h∂ x

+g (s0−s f )=0 (3 )

g(s¿¿0−s f )=0¿ (4 )

The diffusive wave approximation (eqn 3) conserves momentum by balancing the water pressure

gradient, acceleration due to gravity, and resistance due to friction along the boundaries of the channel.

By including the water pressure gradient, the diffusive wave equation allows the energy of the wave to

diffuse, via lengthening and flattening, as the wave propagates downstream (Novak, et al. 2010). The

pressure gradient also gives diffusive wave models some ability to account for backwater effects. Since it

neglects terms for local and convective acceleration, the diffusive wave approximation is best suited to

slow flows on gentle slopes without rapid changes in flow (Novak, et al. 2010). The kinematic wave

approximation (eqn 4) assumes that the wave is long and flat enough that the pressure gradient and local

and convective acceleration are negligible in comparison to gravitational acceleration along the channel

slope (Miller 1984). Kinematic waves propagate downstream only and thus cannot represent backwater

14

249

250

251

252

253

254

255

256

257

258

259

260

261

262

263

264

265

266

267

268

269

270

Page 15: opus.bath.ac.ukopus.bath.ac.uk/53573/1/Collender_disease_risks_inla… · Web viewopus.bath.ac.uk

effects; these waves are best suited for representation of shallow flows over steep slopes (Novak, et al.

2010) in the absence of highly super-critical flows (Miller 1984).

Some models further simplify flood-wave routing by using storage routing (also termed hydrological

routing), which considers only conservation of mass. By neglecting the conservation of momentum

altogether, storage routing relates inflows, outflows, and volumes of water in a channel reach directly;

see, for example, the flow routing equation (eqn 5) implemented in WATFLOOD (Kouwen 2014):

I 1+ I 2

2−

O1+O2

2=

S2−S1

∆ t(5 )

where I , O, and S are the inflow and outflow and storage (volume) of water in the channel; subscripts 1

and 2 denote quantities at the beginning and end of time interval ∆ t , respectively. In order to determine

I 2 and O2, Q may be estimated from rating curves (eqn 6) or by employing methods such as the

Muskingum equation (eqn 7) which relates the movement of a flood wave to conceptual storages for

steady (prism) and transient (wedge) flows.

Q=α hβ (6 )

Q=θI +(1−θ )O= 1K s

∗S (7 )

where α and β are derived constants; θ is a weighting factor representing the relative effects of inflows

and outflows on storage within the reach; and K s is the ratio of storage to discharge and approximates the

travel time of the flood wave through the reach (Martin and McCutcheon 1998). The Muskingum method

can be shown to be a discretization of the kinematic wave equation. The related Muskingum-Cunge

method, in which numerical dispersion is controlled by channel hydraulic characteristics (Miller 1984),

15

271

272

273

274

275

276

277

278

279

280

281

282

283

284

285

286

287

288

289

290

Page 16: opus.bath.ac.ukopus.bath.ac.uk/53573/1/Collender_disease_risks_inla… · Web viewopus.bath.ac.uk

can be shown to be a discretization of the diffusive wave equation (Koussis 2009). Importantly, storage

routing methods provide no information on variability of flow between the considered inflow and outflow

locations (Koussis 2009). The number of spatial dimensions in any hydrodynamic equation may affect

estimates of flow depth and velocity. For instance, secondary lateral flows assumed to be negligible in 1D

models are accounted for in 2D formulations; hence, differences in the spatial distribution of flows

simulated by either approach result in different simulated flow depths and velocities.

Momentum losses to friction and turbulence often vary dramatically across flow domains, and

implementations of hydrodynamic models may vary in their spatial discretization of roughness

parameters. The problem of calibrating roughness across the spatial domain of a hydrodynamic model is

analogous to issues with subsurface parameters in rainfall-runoff models, in that these parameters are not

directly measurable at the scales used in models. Similar to the case of distributed rainfall-runoff models,

while it seems intuitive that allowing roughness to vary at the scale of spatial discretization would offer

more realistic simulations, observational data are rarely, if ever, capable of identifying roughness

parameters at fine spatial resolutions. Hence, many modelers simply apply distinct roughness parameter

values to floodplains and channels (Beven 2007). Hydrodynamic model approaches to simulating flow

are summarized in Table 2.

Table 2: Hydrodynamic model approaches to routing flow through channels and

floodplains

Model Routing in channel

Channel dimensionality

Routing in floodplain

Floodplain dimensionality References

DIVAST Dynamic wave 2D Dynamic wave 2D(Falconer, et al. 2001, Gao, et al.

2011)

SOBEK/D-Water Quality Dynamic wave 1D Dynamic wave 2D (Deltares

Systems 2013a)

HSPF Storage routing 1D NA NA (Bicknell, et al. 1996)

WATFLOOD Storage routing 1D Storage routing with simplified

floodplain

1D (Wu, et al. 2009, Kouwen

2014, Dorner, et

16

291

292

293

294

295

296

297

298

299

300

301

302

303

304

305

306

307

308

309

Page 17: opus.bath.ac.ukopus.bath.ac.uk/53573/1/Collender_disease_risks_inla… · Web viewopus.bath.ac.uk

geometry al. 2006)

SWATStorage routing,

Muskingum routing

1D

Storage routing, Muskingum routing

with simplified floodplain geometry

1D (Neitsch, et al. 2011)

KazamaET AL. Dynamic wave 1D

Dynamic wave without convective acceleration term

2D(Kazama, et al. 2012, Kazama,

et al. 2007)

Yakirevich et al. Dynamic wave 1D NA NA (Yakirevich, et al. 2013)

QUAL2K Storage routing 1D NA NA (Chapra, et al. 2012)

WAMVIEW Storage routing 1D NA NA

(Soil & Water Engineering

Technology Inc. , Tian, et al.

2002)

FASTER Dynamic wave 1D NA NA

(Kashefipour and Falconer

2002, Falconer and

Kashefipour 2001)

EFDC Dynamic Wave 1D, 2D, or 3D NA NA(Bai and Lung

2005, Tetra Tech Inc. 2002)

DUFLOW Dynamic wave 1D Dynamic wave 1D(1995,

Manache, et al. 2007)

TELEMAC Dynamic wave 2D or 3D Dynamic wave 2D or 3D(Desombre

2013, Bedri, et al. 2011)

SLIM Dynamic wave 1D NA NA

(de Brye, et al. 2010, de

Brauwere, et al. 2011)

MOBED Dynamic wave 1D NA NA(Krishnappan 1985, Droppo,

et al. 2011)

QUAL2E-GIBSI Storage routing 1D NA NA

(Simon, et al. 2013, Brown and Barnwell

1987, Rousseau, et al. 2000)

HEMAT Dynamic wave 2D Dynamic wave 2D

(Namin, et al. 2002,

Schnauder, et al. 2007)

KINEROS2/STWIR Kinematic wave 1D Kinematic wave 1D (2000, Guber, et

al. 2014)

17

Page 18: opus.bath.ac.ukopus.bath.ac.uk/53573/1/Collender_disease_risks_inla… · Web viewopus.bath.ac.uk

WEPP NA NA Kinematic wave 1D

(Flanagan and Nearing 1995,

Yeghiazarian, et al. 2006,

Bhattarai, et al. 2011)

3.2 Exchange of flow between channels and floodplains

During flood events, overbank flows may result in the development of shear layers between slow-moving,

shallow floodplain flows and swifter, deeper channel flows. The transfer of momentum across these shear

layers affects flow velocity and is potentially important for understanding the transport of sediments and

associated pathogens within channels (Ikeda and McEwan 2009). Furthermore, the large-scale turbulent

structures that may result contribute to substantial dispersion of transported constituents into inundated

areas (Besio, et al. 2012). In most compound channel formulations, transfer of momentum between the

channel and floodplain is neglected (Seckin, et al. 2009). Channel-floodplain exchanges in coupled 1D-

2D models are most often estimated using weir or friction slope equations which similarly neglect the

transfer of momentum. Such approaches may be reasonably accurate when there is significant obstruction

(e.g., embankments, levees, etc.) of flow between channels and floodplains. However, for most other

situations numerical 1D-2D coupling techniques conserving both mass and momentum have been shown

to greatly improve the accuracy of modeled velocity fields in floodplains (Bladé, et al. 2012). Methods for

simulating exchanges of flow between channels and floodplains in models that simulate overbank flows

are summarized in Table 3.

Table 3: Hydrodynamic model approaches to simulating exchange of flow between

channels and floodplains

Model Method of exchange between channel and floodplain

Quantities conserved References

DIVAST Floodplains and tidal flats may be modeled on the same finite element mesh as channels. Mass, momentum (Falconer, et al. 2001)

SOBEK/D-Water Quality

Flow above elevation of 2D overland grid cells containing channel elements, or above elevation

Mass (Deltares Systems 2013a)

18

310

311

312

313

314

315

316

317

318

319

320

321

322

323

324

325

326

327

328

Page 19: opus.bath.ac.ukopus.bath.ac.uk/53573/1/Collender_disease_risks_inla… · Web viewopus.bath.ac.uk

of defined embankments enters 2D floodplain model.

WATFLOODCompound channel. Channel and overbank flow

velocities may be affected by separate terrain roughness parameters.

Mass(Wu, et al. 2009,

Kouwen 2014, Dorner, et al. 2006)

SWAT Compound channel. Mass (Neitsch, et al. 2011)

Kazama et al. Weir equations for channel overflow across levees. Mass (Kazama, et al. 2012,

Kazama, et al. 2007)

DUFLOW Compound channel. Mass (1995, Manache, et al. 2007)

TELEMAC Floodplains and tidal flats may be modeled on the same finite element mesh as channels. Mass, momentum (Desombre 2013,

Bedri, et al. 2011)

HEMAT Floodplains and tidal flats may be modeled on the same finite element mesh as channels. Mass, momentum (Namin, et al. 2002,

Schnauder, et al. 2007)

KINEROS2/ STWIRCompound channel. Channel and overbank flow

velocities may be affected by separate terrain roughness, losses to infiltration, and bed slopes.

Mass (2000, Guber, et al. 2014)

4. Modeling pathogen transport during floods

Pathogen transport during flood events can be conceptualized as the result of three general processes:

mobilization of pathogens from sources on the land surface or within channels; transportation of

pathogens within runoff and streamflow; and removal of pathogens from runoff and streamflow via

settling, or die-off. Variations in depth and velocity of flow across flood-affected areas can have a

considerable influence on microbial transport, including on pathogen settling and mobilization, mixing by

mechanical dispersion, and, through the suspension and deposition of solids, die-off mediated by solar

radiation.

19

329

330

331

332

333

334

335

336

337

338

Page 20: opus.bath.ac.ukopus.bath.ac.uk/53573/1/Collender_disease_risks_inla… · Web viewopus.bath.ac.uk

4.1 Source characterization

In order to estimate influxes of pathogens to channels during rainfall events, including those that give rise

to flood conditions, modelers must characterize point sources of pathogens, such as wastewater treatment

plants (WWTPs), and non-point sources, such as previously contaminated channel sediments and diffuse

fecal contamination on the land surface, in their study area. Assessment of sanitation conditions and

sanitary infrastructure for source characterization alone may provide important information regarding

potential microbial risks during flooding (Chaturongkasumrit, et al. 2013, Funari, et al. 2012), including

whether high flows are likely to increase pathogen concentrations in local water, as would be expected in

the event of WWTP or sewer failure (see e.g., Bhavnani, et al. 2014, Baqir, et al. 2012, ten Veldhuis, et

al. 2010, Massoud, et al. 2009, Shimi, et al. 2010); or likely to dilute and flush contamination present in

waterways (see e.g., Bhavnani, et al. 2014).

20

339

340

341

342

343

344

345

346

347

348

349

350

Page 21: opus.bath.ac.ukopus.bath.ac.uk/53573/1/Collender_disease_risks_inla… · Web viewopus.bath.ac.uk

For the most part, pathogen source characterization for flood conditions should follow characterization

approaches for typical rainfall-runoff modeling, which have been discussed in detail elsewhere in the

literature (e.g., de Brauwere, et al. 2014, Benham, et al. 2006, Blaustein, et al. 2015). However, it is worth

noting that the efficiency of microbial release from fecal deposits may change with rainfall intensity and

amount, especially for feces with higher liquid content (Blaustein, et al. 2015). Furthermore, saturation of

soils with water and increases in suspended organic matter will tend to decrease retention of microbes in

soil and sediment matrices, adding to effective mobilization in large-scale models (Bradford, et al. 2013).

Therefore, care should be taken to ascertain whether empirical models for microbial release and transport

derived under nominal conditions are suitable for extreme rainfall and floods. Additionally, key sources,

such as channel bed sediments that provide favorable environments for pathogen survival and may

become resuspended in high flows, and flooded latrines, may be significantly more important under flood

conditions than they are under nominal conditions (Wu, et al. 2009, Bhavnani, et al. 2014, Carlton, et al.

2014, Hofstra 2011, Muirhead, et al. 2004, Jamieson, et al. 2004, Coffey, et al. 2010). The effects of

flooding on pathogen sources may also be socially or behaviorally mediated, as in the Ganges Basin in

Bangladesh, where only 3% of toilets and sanitation remain in use during annual floods while a majority

of the population resorts to defecation in hanging latrines that open directly into the environment (Shimi,

et al. 2010).

4.2 Pathogen mobilization

Subsection 4.2 nomenclatureI pat hk

Influx of pathogens to channel reach from source k(organisms)

Nwwtp Population connected to wastewater treatment plant(individuals)

Pi24Number of pathogens excreted per person per time interval(organisms ∙ individuals-1 ∙ time-1)

V exc Volume of effluent exceeding capacity of wastewater treatment plant(L)

∆ t e Rainfall event duration(units range from seconds to days)

V cap Capacity of wastewater treatment plant prior to overflow(L)

21

351

352

353

354

355

356

357

358

359

360

361

362

363

364

365

366

367

368

369

Page 22: opus.bath.ac.ukopus.bath.ac.uk/53573/1/Collender_disease_risks_inla… · Web viewopus.bath.ac.uk

Mobilization of pathogens during flood events can occur via: 1) erosion of fecal matter or contaminated

soil and sediment from the land surface or from channel boundaries; 2) passive diffusion of unattached

pathogens; and 3) direct release of already contaminated water into the environment (e.g., in the case of

WWTP overflows). In many models, pathogen mobilization is presented as analogous to or dependent on

erosion of soil and sediments as a function of the energy available in rainfall, runoff or streamflow. The

quantity of pathogens mobilized in such a fashion may be expressed using physically-based, mechanistic

equations (presented in section 4.4), or simpler empirical or conceptual relationships (presented in section

4.7).

While many models allow input of time-series data characterizing pathogen concentrations in WWTP

effluent, models describing dependence of pathogen mobilization from sanitation facilities on rainfall or

discharge are rare. The sole example identified in the present review was a transport model coupled to

IHACRES that employs a simple expression for discharges of pathogens from wastewater treatment

plants in wet weather (eqn 8) (Ferguson, et al. 2007):

I pat hwwtp=

Nwwtp Pi24V exc ∆ t e

V exc+V cap(8 )

where Nwwtp is the proportion of the population of a subcatchment connected to a WWTP; Pi24 is the

number of microorganisms excreted per person per day; V exc is the volume of effluent exceeding the

capacity of the WWTP during a wet weather event; ∆ t e is the event duration; and V cap is the treatment

capacity of the WWTP before it overflows. While flow parameters play no explicit role in eqn 8, the

concept of capacity exceedance implicit in eqn 8 may be a reasonable way to approach mobilization of

pathogens from inundated WWTPs. A similar model could relate flow volume over spatial elements

22

370

371

372

373

374

375

376

377

378

379

380

381

382

383

384

385

386

387

388

389

390

391

392

Page 23: opus.bath.ac.ukopus.bath.ac.uk/53573/1/Collender_disease_risks_inla… · Web viewopus.bath.ac.uk

containing rudimentary sanitation infrastructure (e.g., pit latrines) to the release of their microbial

contents; this would be a more relevant approach for estimating pathogen mobilization in many

developing country settings.

4.3 Modeling pathogen transport in flowing water

Subsection 4.3 nomenclatureC path Concentration of pathogens in the water

column(organisms ∙ L-1)

t Time(units range from seconds to days)

v Flow velocity(m ∙ time-1)

x Distance along primary axis of flow(m)

D Diffusion/Dispersion coefficient(m2 ∙ time-1)

G General term for sinks and sources of substances(organisms ∙ time-1 ∙ L-1 or g ∙ time-1 ∙ L-1)

Epath Exported load of pathogens(organisms)

I pat hkInflux of pathogens to channel reach from source k(organisms)

δ 24 Fraction of pathogens surviving after 24 hours

LR lLocal reach length for transport to main channel (square root of sub-catchment area)(m)

Πd Probability of pathogen settling out of flow over 1 km reach

Hydrodynamic models typically simulate pathogen transport processes using the advection-diffusion

equation (ADE; eqn 9) which models the transport of constituents based on conservation of mass:

∂C path

∂ t=−v

∂ C path

∂x+D

∂2C path

∂ x2 +G (9 )

where C pathis the concentration of transported pathogens; v is flow velocity; D is a coefficient describing

the magnitude of passive diffusion, turbulent diffusion, or dispersion. The first term on the right side of

the equation, −v∂ Cpath

∂ x, represents advection of the substance in the direction of flow, the second term,

23

393

394

395

396

397

398

399

400

401

402

403

404

Page 24: opus.bath.ac.ukopus.bath.ac.uk/53573/1/Collender_disease_risks_inla… · Web viewopus.bath.ac.uk

D∂2C path

∂ x2 , represents diffusion or dispersion of the substance to areas of lower concentration, and the

final term, G, is a composite term representing source processes adding pathogens to the water column

(e.g., influx in runoff or from upstream channels, suspension from bed sediments, and growth) as well as

sink processes removing pathogens from the water column (e.g., sedimentation along the channel bed and

pathogen die-off). Transport of sediment-associated and free-floating pathogens in the water column may

be modeled separately by using different values for the diffusion coefficient (Jamieson, et al. 2005). The

ADE has been noted to underestimate the spreading of tracer concentrations in many natural rivers, and

attribute this shortcoming to the assumption within the equation that the diffusive or dispersive flux is

proportional to the local concentration of transported material (Beven 2007, Blazkova, et al. 2012).

Instead, transport time distributions in many river systems are dominated by the effects of retention in

areas of low flow velocity near channel boundaries, so-called ‘dead zones’, which depend primarily on

turbulence structures developing from the geometry of the flow domain and the volume of flow (Beven

2007, Blazkova, et al. 2012). The ADE can be augmented to describe these effects (see e.g. (Bencala and

Walters 1983, Romanowicz, et al.)), but doing so requires the addition of local parameters that must be

calibrated and may exacerbate problems of identifiability and equifinality (Blazkova, et al. 2012).

Alternatively, Blazkova et al. (2012) (Blazkova, et al. 2012) proposed using a simple linear transfer

function, originally suggested by Beer and Young (1983) (Beer and Young 1983), to capture the effect of

dead-zone retention on the transport time distribution of suspended constituents, although this approach

has not been validated for flood conditions.

Simplified transport models may neglect the diffusion term or may use mixing cell techniques. The

relationship of mixing cell techniques to the advection-diffusion equation is analogous to the relationship

between the Muskingum wave routing technique to the kinematic wave shallow-water equations, in that

they may be regarded as finite-difference numerical solutions to the advection-diffusion equation (Barry

and Bajracharya 1996). However, dispersion/diffusion processes are not explicit in mixing cell models.

24

405

406

407

408

409

410

411

412

413

414

415

416

417

418

419

420

421

422

423

424

425

426

427

428

429

Page 25: opus.bath.ac.ukopus.bath.ac.uk/53573/1/Collender_disease_risks_inla… · Web viewopus.bath.ac.uk

Instead, the mixing cell models that attempt to represent dispersion/diffusion do so by controlling

numerical dispersion through varying the size of spatial and temporal intervals for calculation (Barry and

Bajracharya 1996). All mixing cell approaches assume pathogen concentrations to be homogenous

throughout a channel reach. Hydrodynamic model approaches to simulating pathogen transport in

channels and floodplains are summarized in Table 4.

Table 4: Hydrodynamic model approaches to pathogen transport

Hydrodynamic model Approach to pathogen transport References

DIVAST 2D advection-diffusion equationSeparate terms for particle-associated and free-floating pathogens (Gao, et al. 2011)

SOBEK/D-Water Quality 1D advection-diffusion equation (Deltares Systems 2013b)

HSPF 1D mixing cell approachSeparate terms for particle-associated and free-floating pathogens (Bicknell, et al. 1996)

WATFLOOD/ DORNER/ WU

1D mixing cell approachSeparate terms for particle-associated and free-floating pathogens

(Wu, et al. 2009, Dorner, et al. 2006)

Kazama et al. 2D advection equation with numerical diffusion

(Kazama, et al. 2012)(S. Kazama,

personal communication,Feb. 3, 2016)

YAKIREVICH ET AL.

1D advection-diffusion equationTerms for exchange of pathogens between channel flow, groundwater,

and transient storage (stagnant pools, eddies, etc.)(Yakirevich, et al. 2013)

QUAL2K 1D mixing cell approach (Chapra, et al. 2012)

WAMVIEW 1D mixing cell approach (Tian, et al. 2002)

FASTER 1D advection-diffusion equation (Falconer and Kashefipour 2001)

EFDC 3D advection-diffusion equationSeparate terms for particle-associated and free-floating pathogens (Bai and Lung 2005)

DUFLOW 1D advection-diffusion equation (Manache, et al. 2007)

TELEMAC 2D advection-diffusion equation (Bedri, et al. 2011)

SLIM 1D advection-diffusion equation (de Brye, et al. 2010, de Brauwere, et al. 2011)

MOBED 1D mixing cell approachExplicitly models particle-associated pathogens (Droppo, et al. 2011)

QUAL2E-GIBSI 1D advection-dffusion equation(Simon, et al. 2013, Brown and Barnwell

1987)

KINEROS2 / STWIR 1D advection-diffusion equation (Guber, et al. 2014)

HEMAT 2D advection-diffusion equation (Schnauder, et al. 2007)

25

430

431

432

433

434

435

Page 26: opus.bath.ac.ukopus.bath.ac.uk/53573/1/Collender_disease_risks_inla… · Web viewopus.bath.ac.uk

In addition to the approaches outlined above, it is worth noting the possibility of augmenting semi-

distributed rainfall-runoff models with simplified conceptual pathogen transport. For instance, Ferguson

et al. (2007) modeled in-channel transport alongside a semi-distributed implementation of IHACRES by

expressing pathogen export for each sub-catchment as a function of steady flow velocity and static

probabilities of pathogen removal by sedimentation or inactivation (eqn 10):

Epath=∑k=1

Nk

I pat hkδ 24

LR l

v ( 1−Π d )LR l (10 )

where Epath is the exported load of pathogens from the sub-catchment; I pat hk is the input of pathogens to

the stream from source k; δ 24 is the fraction of pathogens surviving in water after 24 hours; LR l is the

local reach length (for transport to the main channel, assumed equal to square root of sub-catchment

area); v is flow velocity over the reach; and Π d is the probability of pathogens settling out over a 1-km

reach. Unfortunately, the authors did not compare model estimates with measured pathogen

concentrations, and the extent to which this approach is capable of representing pathogen transport under

real conditions is not readily apparent.

4.4 Hydrodynamic calculation of pathogen suspension and settling during floods

Section 4.4 nomenclatureGmob Net source or sink of substance to water

column from deposition and suspension(organisms ∙ time-1 ∙ L-1 or g ∙ time-1 ∙ L-1)

Rrl Rate of suspension of substance l(organisms ∙ time-1 ∙ L-1 or g ∙ time-1 ∙ L-1)

Rd l Rate of deposition of substance l(organisms ∙ time-1 ∙ L-1 or g ∙ time-1 ∙ L-1)

ce Entrainment coefficient(g ∙ m-2 ∙ time-1)

τ b Shear stress along bed(Pa)

τ crCritical shear stress for net suspension(Pa)

τ cdCritical shear stress for net deposition(Pa)

ω Particle settling velocity(m ∙ time-1)

26

436437438

439

440

441

442

443

444

445

446

447

448

449

450

451

452

453

454

Page 27: opus.bath.ac.ukopus.bath.ac.uk/53573/1/Collender_disease_risks_inla… · Web viewopus.bath.ac.uk

Psus Number of pathogens suspended(organisms)

C pathsed Concentration of pathogens in bed sediments(organisms ∙ g-1)

C sed maxMaximum concentration of sediment channel can transport(g ∙ L-1)

C sed Concentration of sediment in channel(g ∙ L-1)

cse Channel erodibility factor(unitless)

cveg Vegetative cover factor(unitless)

Sch Storage (volume) of water in channel(L)

Pdep Number of pathogens deposited(organisms)

t Time(units range from seconds to days)

t cr Time at which critical shear stress for suspension is exceeded(time)

K e Enhanced mass transfer rate from sediments to water after biofilm sloughing(g ∙ time-1)

ρ sed Bed sediment density(g ∙ L-1)

ρ Water density(g ∙ L-1)

g Acceleration due to gravity(m ∙ time-2)

vx , v y Flow velocity in x and y dimensions(m ∙ time-1)

C Chezy roughness coefficient(m1/2 ∙ time-1)

τ c Pivot point critical shear stress -suspension dominates above, deposition dominates below(Pa)

s0 Channel bed slope(unitless)

R Hydraulic radius(m)

Q Discharge (rate of flow)(L ∙ time-1)

α General term for user-defined or calibrated coefficient

vpk Peak flow velocity(m ∙ time-1)

β General term for user-defined or calibrated exponent

t res Residence time in reach(time)

W Channel width(m)

W b Bottom width of channel(m)

v Mean flow velocity(m ∙ time-1)

V c h24Daily flow volume(L)

V c h24cr Critical daily flow volume determining dominance of suspension or deposition(L)

mme Dry sediment mass per unit area(g ∙ m-2)

∆ t Length of time step(time)

Av Vertical turbulent viscosity(Pa ∙ time)

z Vertical coordinate within water column(m)

sf Friction slope (energy head loss)(unitless)

27

Page 28: opus.bath.ac.ukopus.bath.ac.uk/53573/1/Collender_disease_risks_inla… · Web viewopus.bath.ac.uk

Net mobilization of pathogens (i.e., the total number entering the water column less the number leaving

through settling or retention) from sources in channels and on land surfaces (included within the G term

in eqn 9) depends on flow velocity, which determines the amount of energy available to mobilize

particles, and flow depth, which influences the time it takes suspended particles to settle. The physical

mechanisms involved include entrainment and settling of particles mediated by shear stresses along the

boundaries of flow, as well as loss of suspended pathogens infiltrating into the soil column or channel

bed. Additionally, pathogens trapped in sediment or soil near the surface-subsurface interface may be re-

mobilized during high flows independent from sediment suspension (Ghimire and Deng 2013,

Yakirevich, et al. 2013, Grant, et al. 2011); however, this hyporheic transport process is rarely accounted

for in fate and transport models (Piorkowski, et al. 2014).

Mobilization of channel bed sediments is an important source of suspended pathogens during high flow

events (Wu, et al. 2009). Physically based models often calculate net mobilization of particle-associated

pathogens as a piecewise function of bed shear stresses (e.g., eqns 11-13; Yakirevich, et al. 2013):

Gmob=Rr path−Rdpath (11)

Rr path={ce( τb

τ c r

−1) for τ b> τcr

0 for τ b≤ τ cr

(12 )

Rd path={ω (1−τb

τ cd) for τb< τcd

0 for τb ≥ τcd

(13 )

28

455

456

457

458

459

460

461

462

463

464

465

466

467

468

469

470

471

472

Page 29: opus.bath.ac.ukopus.bath.ac.uk/53573/1/Collender_disease_risks_inla… · Web viewopus.bath.ac.uk

where Gmob is the net flux of pathogens into or out of the water column via suspension and sedimentation;

Rr path is the rate of pathogen resuspension; Rd path is the rate of pathogen deposition; ce is an entrainment coefficient; τ b is the shear stress along the channel bed; τ cr

and τ cd are empirically

determined critical shear stresses for erosion and deposition, respectively; and ω is the particle settling

velocity (calculated via Stokes’ law or other means). Depending on each model’s approach to

hydrodynamics, turbulent shear stresses may be calculated as a function of either temporally and spatially

varying quantities, such as flow depth and velocity, or spatially varying but temporally invariant factors

such as channel slope and width, hydraulic radius, and friction coefficients. Critical shear stresses for

erosion or deposition may be empirically determined or estimated based on sediment properties.

Less physically oriented models for pathogen mobilization and deposition rely, respectively, on power

laws relating suspension to discharge or flow velocity, or expressions relating settling rates to particle

settling velocity, which may or may not incorporate flow depth or pathogen concentrations (Wu, et al.

2009, de Brauwere, et al. 2011, de Brye, et al. 2010, Chapra, et al. 2012). Additionally, many models

which implement storage routing invoke a concept of sediment transport capacity. Settling and

suspension are specified as mutually exclusive conditions resulting when the concentration of transported

sediment is above or below the channel’s transport capacity, respectively (e.g., eqn 14; Neitsch, et al.

2011), or in a roughly equivalent specification, the critical daily flow volume for erosion (Tian, et al.

2002):

{Psus=Cpat h sed∗(C se dmax

−C sed )∗cse∗c veg∗Sch

for C sed<C sed max

Pdep=(C sed−C se dmax )∗Sch

for C sed>C sed max

(14 )

29

473

474

475

476

477

478

479

480

481

482

483

484

485

486

487

488

489

490

491

492

493

Page 30: opus.bath.ac.ukopus.bath.ac.uk/53573/1/Collender_disease_risks_inla… · Web viewopus.bath.ac.uk

where Psus is the number of pathogens suspended; C sed max is the capacity of the channel reach to transport

sediment (which may be determined as a function of peak flow velocity, average flow velocity, depth of

flow, channel slope, or shear stresses); C pat hsed is the concentration of pathogens in bed sediments; C sed is

the concentration of sediment currently transported by the channel; cse is a coefficient representing the

erodibility of the bed material; cveg is a coefficient representing the effect of vegetation on bed erodibility;

Sch is the storage (volume) of water in the channel; and Pdep is the number of pathogens deposited.

Some investigators have reported evidence that sloughing of biofilms during high flows may contribute to

enhanced mobilization of pathogens from bed sediments (Yakirevich, et al. 2013, Droppo, et al. 2007,

Droppo, et al. 2009). To date, this mechanism has yet to be incorporated into a fate and transport model,

though a piecewise function has been proposed (eqn 15; Yakirevich, et al. 2013):

Psus (t> tcr )=Rr sedC pat hsed

+ K e ρ sed Cpat hsedH ( t−tcr ) (15 )

where Rr sed is the rate of resuspension of sediment prior to sloughing of biofilms; C pat hsedis the

concentration of bacteria in bed sediments; K e is the mass transfer rate due to enhanced erosive exchange;

ρ sedis the density of bed sediments; H (t ) is the Heaviside step function; and t cr is the time at which

critical bed shear stress for suspension is exceeded (Yakirevich, et al. 2013). Table 5 summarizes

hydrodynamic model approaches for calculating net pathogen mobilization from bed sediments.

Table 5: Hydrodynamic model approaches to net pathogen mobilization from bed sediments

Hydrodynamic model General approach Suspension determinants

Settling determinants

Shear stress / capacity

formulation

References

DIVASTPiecewise functions

of shear stressesτ b , τ cr

ω ,C sed , τb , τ c ❑d

τ b=¿

ρg ( v x+v y )C2

(Falconer, et al. 2001,

Gao, et al. 2011)

30

494

495

496

497

498

499

500

501

502

503

504

505

506

507

508

509

510

511

512

Page 31: opus.bath.ac.ukopus.bath.ac.uk/53573/1/Collender_disease_risks_inla… · Web viewopus.bath.ac.uk

HSPFMutually exclusive piecewise functions

of shear stressescse , τb , τc ω ,C sed , τb , τc τ b=s0 ρR

(Bicknell, et al. 1996)

WATFLOOD/ WU

Power law for suspension, deposition

dependent only on settling velocity

Q ω NA (Wu, et al. 2009)

SWAT

(OPTION 1)

Mutually exclusive piecewise functions of sediment carrying

capacity

C sed max, C sed , cse ,cveg

C sed max,C sed C sed max

=α∗v pkβ (Neitsch,

et al. 2011)

SWAT

(OPTIONS 2-5)

Piecewise function of sediment carrying capacity, calculated from various stream power formulations,

with constant deposition. Erosion

is calculated separately for banks

and bed.

cse , τb , τ c , ρb ,C sed max

, C sedω ,t res , h

τ b=ρh s0( W2W b

+0.5)(Neitsch, et al. 2011)

Yakirevich et al.Piecewise functions

of shear stressesτ b, τ cr s

ω ,τb , τ cr d τ b=ρ cd v2(Yakirevich, et al.

2013)

QUAL2KNo suspension,

simple deposition NA ω, h NA(Chapra,

et al. 2012)

WAMVIEW

Mutually exclusive piecewise functions

of daily flow volume

V c h24,V c h24cr

V c h24,V c h24cr

NA (Tian, et al. 2002)

EFDCPiecewise functions

of shear stresses

Mode 1: Gradual Erosion

ρb , τb , τc ❑r

Mode 2: Mass Erosion

ρb , τb , mme ,∆ t

ω ,C sed , τb , τcd τ b=

Av

h∗dv

dz

(Bai and Lung 2005, Tetra

Tech Inc. 2002)

SLIMNo suspension,

simple deposition NA C path, ω, h NA

(de Brye, et al.

2010, de Brauwere

, et al. 2011)

MOBED

Non-exclusive piecewise functions for suspension and

deposition, dependent on shear

stresses

τ b , τ c τ b , τ c τ b=ρgR sf

(Droppo, et al. 2011,

Krishnappan

1981)

31

513

Page 32: opus.bath.ac.ukopus.bath.ac.uk/53573/1/Collender_disease_risks_inla… · Web viewopus.bath.ac.uk

4.5 Transport via mechanical dispersion during floods

Section 4.5 nomenclatureD Dispersion coefficient

(m2 ∙ time-1)h Depth of flow

(m)v¿

Shear velocity - √ τb

ρ(m ∙ time-1)

v Flow velocity(m ∙ time-1)

W Channel width(m)

g Acceleration due to gravity(m ∙ time-2)

cd Dispersivity or diffusion constant(m2 ∙ time-1)

n Manning roughness coefficient(time ∙ m-1/3)

v Mean flow velocity(m ∙ time-1)

Variations in the velocity profile along horizontal and vertical axes of flow, such as those that develop

between channels and inundated areas during flooding, result in mechanical dispersion, in which uneven

flow velocities mix and distribute transported constituents along the flow’s longitudinal and/or transverse

axes (Fischer, et al. 1979). The effect of mechanical dispersion is calculated within the advection-

diffusion equation (eqn 8) through the parameter D. In 1D formulations, D represents longitudinal

dispersion along the axis of flow while, in 2D formulations, separate variables account for longitudinal

dispersion and transverse dispersion. The values of the dispersion coefficients are affected by a large

number of flow and channel geometry parameters and several investigators have proposed empirical

formulae to relate dispersion to other quantities. These quantities typically include various combinations

of flow and shear velocities, channel width, and water depth, e.g., as in models proposed by Elder (1959)

(eqn 16), Fischer (1975) (eqn 17), and Kashefipour and Falconer (2002) (eqn 18):

D=5.93 h v¿

(16 )

D=0.011 v2W 2

hv¿(17 )

32

514

515

516

517

518

519

520

521

522

523

524

525

526

527

Page 33: opus.bath.ac.ukopus.bath.ac.uk/53573/1/Collender_disease_risks_inla… · Web viewopus.bath.ac.uk

D=(7.428+1.775( Wh )

0.620

( v¿

v )0.572)hv ( v

v¿ ) (18 )

where W is channel width; v is flow velocity along the primary axis of flow;v¿ is shear velocity; and h is flow depth. The model proposed by Elder (1959) follows from assumptions of a logarithmic velocity profile across the depth of an infinitely wide open channel.

However, in actual channels, the effect of the lateral shear velocity profile between the two banks can

increase D by up to three orders of magnitude (Fischer, et al. 1979). Thus, the models developed by

Fischer (1975) and Kashefipour and Falconer (2002) incorporate channel width to account for the effects

of friction along the banks. Following the publication of Kashefipour and Falconer’s model (2002), which

was derived empirically, another group of authors (Toprak, et al. 2004) raised questions about the validity

of some of the methods underlying the model. Nonetheless, eqn 18 has been found to outperform other

simple formulations for longitudinal dispersion, though neural network models have been used to arrive at

even more accurate predictions (Toprak and Cigizoglu 2008). In more complex 2D and 3D models,

dispersion may be more accurately accounted for by explicit representation of secondary currents and

turbulence propagation (Wallis and Manson 2005). The approaches of existing microbial fate and

transport models for characterizing dispersion are summarized in Table 6.

Table 6: Hydrodynamic model approaches to estimating dispersion in channels and floodplains

Hydrodynamic model Approach to pathogen dispersion References

DIVAST

Calculated from flow velocity in 2 dimensions, total water depth, and Chezy roughness coefficient:

D xx=√g h (5.93 vx

2+0.23 v y2 )

C √vx2+v y

2

D xy=D yx=√ g h (5.93−0.23 ) vx v y

C √v x2+v y

2

D yy=√g h (5.93 v y

2 +0.23 vx2 )

C √vx2+v y

2

(Falconer, et al. 2001)

33

528

529

530

531

532

533

534

535

536

537

538

539

540

541

542

543

Page 34: opus.bath.ac.ukopus.bath.ac.uk/53573/1/Collender_disease_risks_inla… · Web viewopus.bath.ac.uk

Yakirevich et al.Calculated from flow velocity:

D=cd v (Yakirevich, et al. 2013)

QUAL2K

Calculated from water depth, channel slope, channel width, and flow velocity using Fischer’s model:

D=0.011∗v2W 2

h v¿

(Chapra, et al. 2012)

FASTER

Calculated from water depth, channel width, flow velocity, and shear velocity using Kashefipour and Falconer model:

D=(7.428+1.775( Wh )

0.62

( v¿

v )0.572)hv ( v

v¿ )(Falconer and Kashefipour

2001)

EFDC Explicit turbulence modeling (Bai and Lung 2005, Tetra Tech Inc. 2002)

TELEMAC Explicit turbulence modeling (Desombre 2013)

QUAL2E-GIBSI

Calculated from roughness, mean flow velocity, and depth of flow. Estimated with Elder Model with customizable dispersion

constant:

D=3.82∗cd∗n∗v∗h56

(Simon, et al. 2013, Brown and Barnwell 1987,

Rousseau, et al. 2000)

4.6 Pathogen persistence during floods

Subsection 4.6 nomenclatureC path Concentration of pathogens in the water

column(organisms ∙ L-1)

t Time(units range from seconds to days)

Kd Pathogen die-off rate constant(time-1)

Ra dsnetNet solar radiation (J ∙ cm-2 ∙ day-1)

R Reflectance due to suspended solids (%) Ra ds¿¿Total downward solar radiation(J ∙ cm-2 ∙ day-1)

C ss Concentration of suspended solids in the water column (mg ∙ L-1)

h Depth of flow (m)

34

544

545

Page 35: opus.bath.ac.ukopus.bath.ac.uk/53573/1/Collender_disease_risks_inla… · Web viewopus.bath.ac.uk

The persistence of pathogens in natural waters is influenced by a variety of factors including temperature,

salinity, pH, competition and predation from other microorganisms, available nutrients, and sunlight

(Hipsey, et al. 2008). During high flows, mobilization of organic matter and sediments, as well as changes

in the depth of the water column, may alter the balance of nutrients available to support populations of

bacterial pathogens, and influence light penetration and ultraviolet radiation in the water column. The net

mobilization of sediments during flood events is, as described in section 4.4, dependent on depth and flow

velocity. Elevated concentrations of mobilized sediments during artificial flood events have been found to

protect pathogens from ultraviolet radiation, resulting in extended persistence of fecal indicator bacteria

by up to 20 hours (Walters, et al. 2014). Mobilized sediments may thus have a substantial impact on

microbiological risks during flood events.

The impact of flood events on nutrients available to microorganisms within the water column varies

depending on the dominant flow paths under base flow and high water conditions, as well as the

distribution of nutrient sources within a catchment. Periods of high flow have been linked to increased in-

channel concentrations of dissolved organic carbon (DOC) in temperate and subtropical climates (Leff

and Meyer 1991, Royer and David 2005, Tesi, et al. 2013), though in some cases the nutritional quality of

DOC was found to decrease (Leff and Meyer 1991). High water was found to result in enhanced bacterial

growth within certain catchments in the Amazon, which is thought to be linked to the influx of

bioavailable nutrients from inundated floodplains (Benner, et al. 1995). Similarly, in many boreal

catchments, flood events linked to snowmelt have been found to increase bioavailable dissolved organic

nitrogen (Stepanauskas, et al. 2000). While increased nutrient availability tends to promote bacterial

persistence and growth, some researchers have observed that the activities of protozoan species grazing

on bacterial populations may also be enhanced under these conditions, though protozoan grazing shows

marked spatial heterogeneity in the environment (Kinner, et al. 1997, Kinner, et al. 1998, Kinner, et al.

2002). Viral persistence may also be influenced through the transport of organic matter, and enhanced by

35

546

547

548

549

550

551

552

553

554

555

556

557

558

559

560

561

562

563

564

565

566

567

568

569

570

571

Page 36: opus.bath.ac.ukopus.bath.ac.uk/53573/1/Collender_disease_risks_inla… · Web viewopus.bath.ac.uk

decreased binding to and inactivation on certain soils and sediments (Blanford, et al. 2005, Pieper, et al.

1997, Ryan, et al. 1999), or diminished through the production of antiviral compounds by microbial

communities flourishing in nutrient-rich conditions (Deng and Cliver 1995). Thus it is difficult to

generalize the effects of flood events on bioavailable nutrient concentrations, as well as the net effect of

increased nutrient availability on bacterial and viral persistence, though this is clearly an area that merits

further study.

Expressions for the inactivation of pathogens within fate and transport models typically take the form of

an exponential function assuming first order kinetic dependency of the rate of inactivation on pathogen

concentrations (eqn 19):

C path( t )=C path (0 )∗e−K d t (19 )

where C path ( t ) is the concentration of pathogens at time t , and K d is a rate constant describing the

geometric change in pathogen concentration over time. The rate constant Kd may be related to factors

impacting pathogen survival, such as temperature or salinity. Effects of suspended solids have rarely been

incorporated into pathogen inactivation functions within existing fate and transport models, while the

effects of nutrient concentrations are (somewhat understandably) completely absent (Table 7). Among the

reviewed models, Kazama et al. (2012) provide the sole model (eqns 20-22) that incorporates mitigating

effects of water depth and suspended solids on the UV-mediated inactivation of pathogens, based on

empirical relationships between reflectance (R) and suspended solid concentration (C ss ) presented by Oki

et al. (Oki, et al. 2001) and empirical exponential relationships between the die-off ratio (e−Kd ¿ and

average net solar radiation at various depths of flow taken from Gameson and Saxon (Gameson and

Saxon 1967). Presumably eqn. 22 was implemented in some continuous fashion within the model used by

Kazama et al., but its exact implementation is not clear from the text of their study.

36

572

573

574

575

576

577

578

579

580

581

582

583

584

585

586

587

588

589

590

591

592

593

594

595

Page 37: opus.bath.ac.ukopus.bath.ac.uk/53573/1/Collender_disease_risks_inla… · Web viewopus.bath.ac.uk

Ra dsnet=1−R

100∗Ra dsmax

(20 )

R=0.0809+0.0146∗C ss (21)

{if h=0.18 m ,e−Kd=117.64∗e−0.0033∗Rad snet

if h=1.00 m,e−Kd=217.33∗e−0.0029∗Rad snet

if h=2.00 m, e−Kd=204.9∗e−0.0021∗Ra dsnet

if h=3.00 m , e−Kd=113.59∗e−0.0010∗Ra d snet

if h=4.00 m ,e−Kd=129.97∗e−0.0007∗Ra ds net

(22)

Table 7: Model approaches to pathogen persistence

Model Inactivation model Rate modifiers Separate rates incorporated References

DIVAST First order decay NA NA (Gao, et al. 2011)

SOBEK/D-WATER QUALITY

First order decay Salinity, temperature, radiation NA (Deltares Systems

2013b)

HSPF First order decay Temperature, radiation (free-floating pathogens)

Free-floating, suspended sediment-associated, and bed sediment-associated

pathogens

(Bicknell, et al. 1996)

WATFLOOD/ DORNER/ WU First order decay NA Winter, spring/fall, summer

months(Wu, et al. 2009,

Dorner, et al. 2006)

SWAT First order decay Temperature

Persistent and less-persistent pathogens

Foliage, soil, soil solution, channel waters.

(Neitsch, et al. 2011)

KAZAMA ET AL. First order decay Solar radiation, water depth, suspended solids NA (Kazama, et al.

2012)

YAKIREVICH ET AL. First order decay NA Pathogens in water column

and bed sediments(Yakirevich, et al.

2013)

QUAL2KFirst order decay, Beer-Lambert law

decay

Temperature, radiation, water depth NA (Chapra, et al.

2012)

WAMVIEW First order decay NA NA (Tian, et al. 2002)

FASTER First order decay NACoastal waters, rivers, dry

weather, wet weather, daytime, nighttime

(Kashefipour and Falconer 2002)

EFDC First order decay Temperature, solar radiation, salinity, pH

Free-floating, suspended sediment-associated, bed

sediment-associated pathogens

(Bai and Lung 2005)

DUFLOW First order decay NA NA (Manache, et al. 2007)

37

596

Page 38: opus.bath.ac.ukopus.bath.ac.uk/53573/1/Collender_disease_risks_inla… · Web viewopus.bath.ac.uk

TELEMAC First order decay NA NA (Bedri, et al. 2011)

SLIM First order decay Temperature NA (de Brauwere, et al. 2011)

HEMAT First order decay NA Daytime, nighttime (Schnauder, et al. 2007)

IHACRES/ FERGUSON First order decay NA

Cryptosporidium, Giardia, E. Coli

Soil, water

(Ferguson, et al. 2007)

KINEROS2/STWIR First order decay NA Applied manure, soil, soil

solution, runoff water (Guber, et al. 2012)

COLI

First order decay between defecation and mobilization in

runoff

Temperature NA (Walker, et al. 1990)

WAMVIEW/ TIAN ET AL. First order decay Temperature, radiation NA (Tian, et al. 2002)

WEPP First order decay NA NA(Yeghiazarian, et al. 2006, Bhattarai, et

al. 2011)

4.7 Conceptual approaches to pathogen transport in rainfall-runoff modeling systems

Subsection 4.7 nomenclatureE Soil loss / sediment yield

(Mg ∙ m-2 ∙ yr-1)crer USLE rainfall erosivity index

(MJ ∙ mm ∙ m-2 ∙ h-1 ∙ yr-1)cserd USLE soil erodibility index

(Mg ∙ m2 ∙ h ∙ m-2 ∙ MJ-1 ∙ mm-1)c LS USLE Length / slope factor

(unitless)ccrm USLE cropping / management factor

(unitless)ccp USLE conservation practice factor

(unitless)α General term for user-defined or

calibrated coefficientV r o24

24 hour runoff volume(L ∙ day-1)

Qmax Peak discharge during storm event(L ∙ time-1)

Als Land surface area contributing to sediment/pathogen transport(m2)

β General term for user defined or calibrated exponent

Pro Number of pathogens transported in runoff(organisms)

Prf Number of pathogens mobilized by raindrop impact(organisms)

Pof Number of pathogens mobilized by overland flow(organisms)

Rrf Rainfall rate (intensity)(mm ∙ time-1)

c gc Hartley model vegetative cover factor (unitless)

ccan Hartley model canopy factor(unitless)

c perd Hartley model pathogen erodibility factor(organisms ∙ J-1)

c ff Hartley model friction parameter(unitless)

ρ Density of water(g ∙ L-1)

sls Overland slope SEDdetrMass of sediment available for

38

597

598

Page 39: opus.bath.ac.ukopus.bath.ac.uk/53573/1/Collender_disease_risks_inla… · Web viewopus.bath.ac.uk

(unitless) transport due to detachment by rainfall(g)

SE Dof Mass of sediment detached and transported in overland flow(g)

cman HSPF supporting management practice factor(unitless)

Sls Water storage on land surface(L)

C sed maxMaximum concentration of sediment runoff can transport(g ∙ L-1)

SE D ro Total sediment mass transported in runoff(g)

cwsh HSPF pathogen washoff factor(mm-1)

While the majority of this review has been devoted to the implications of physical process representation

in hydrodynamic models for modeling pathogen transport during floods, most rainfall-runoff models

represent pathogen mobilization and transport in a conceptual or empirical fashion. In these models, fine-

scale physical processes underlying pathogen transport processes are represented implicitly, if at all.

Watershed-scale pathogen transport modelling is frequently challenging due to lack of site-specific data

on pathogen discharges, the complexity and diversity of pathogen characteristics, and uncertainties

inherent in the underlying environmental models, particularly those related to obtaining averaged, or

effective parameter values at the spatial scale of the model when available theory and measurements

describe these processes at small scale, or in artificially homogeneous settings . Obtaining accurate

estimates of pathogen transport in runoff under extreme rainfall and flooding conditions presents

additional challenges, as common empirical relationships and simplifying assumptions employed by

rainfall-runoff models may not hold.

Characteristics of rainfall and runoff, such as rainfall pattern, effective rainfall volume, and overland flow

velocity, have been identified as some of the most critical factors affecting transport of pathogens from

the land surface (Funari, et al. 2012, Bhavnani, et al. 2014, Jamieson, et al. 2004, Ferguson, et al. 2005,

Jung, et al. 2014, Tsihrintzis and Hamid 1997). However, the relationship of these factors with in-stream

pathogen concentrations across sites is complex. For instance, some investigators have reported dilution

of pathogen concentrations by extreme rainfall (e.g. (Bhavnani, et al. 2014)) while others have noted

39

599

600

601

602

603

604

605

606

607

608

609

610

611

612

613

614

615

616

617

618

Page 40: opus.bath.ac.ukopus.bath.ac.uk/53573/1/Collender_disease_risks_inla… · Web viewopus.bath.ac.uk

pathogen concentrations several orders of magnitude higher after storm events (Bhavnani, et al. 2014,

Page, et al. 2012). These disparate results likely have to do with pathogen sources and drainage

characteristics of the study area. Other factors used to predict pathogen loading during storms or flood

events include land-use factors, such as the presence of manure-fertilized areas and measures to manage

agricultural runoff, catchment topography, and soil and sediment characteristics (Jamieson, et al. 2004,

Ferguson, et al. 2007, Papanicolaou 2008).

Typically, models for pathogen transport from the land surface involve five components: 1) deposition of

pathogens on the land surface from point or non-point sources (e.g. domestic animals, faulty sanitation);

2) accumulation of pathogens in various reservoirs (e.g. soils, subsurface, foliage); 3) removal of

pathogens via die-off or irreversible infiltration into soils; 4) mobilization of pathogens from

environmental reservoirs by erosion, raindrop impacts, or flowing water; and 5) pathogen transport in

overland and subsurface flows rainwater or overland flow (Figure 1). Here, we restrict our discussion to

mobilization and transport in overland flows, as subsurface flows are rarely incorporated into rainfall-

runoff pathogen transport models, and transport in overland flow is expected to dominate during flood

events; we omit further discussion of environmental reservoirs, pathogen die-off, and influxes to the land

surface from pathogen sources, as these have been discussed in previous sections or are not particularly

relevant to flood conditions.

40

619

620

621

622

623

624

625

626

627

628

629

630

631

632

633

634

635

636

Page 41: opus.bath.ac.ukopus.bath.ac.uk/53573/1/Collender_disease_risks_inla… · Web viewopus.bath.ac.uk

Figure 1- General components of rainfall-runoff pathogen transport modules.

In many rainfall-runoff modelling systems, pathogen mobilization and transport is conceptualized as

being analogous to or correlated with the mobilization and transport of sediments from the land surface

brought about by raindrop impacts or overland flow (Kouwen 2014, Ferguson, et al. 2003). Thus, many

expressions for pathogen transport in these models are derived from sediment delivery models such as the

Universal Soil Loss Equation (USLE; eqn 23):

E=crer∗c serd∗cLS∗ccrm∗ccp (23 )

where E is sediment yield / soil loss; crer is a rainfall erosivity index (an empirical expression for the

kinetic energy of rainfall as a function of its intensity (Elbasit, et al. 2011)); cserd is a soil erodibility

index; c LS is a composite factor describing the length and gradient of the hillslope; ccrm is a cropping

management factor; and ccp a supporting conservation practice factor (Merritt, et al. 2003, Aksoy and

Kavvas 2005, Renard, et al. 1991). The Revised Universal Soil Loss Equation (RUSLE), a descendant of

the USLE, features refinements to the estimation of the terms of the USLE that increase the level of

process representation and, therefore, the generalizability (Lane, et al. 1992). In contrast to the USLE and

RUSLE, the Modified Universal Soil Loss Equation (MUSLE) was developed for prediction of sediment

41

Influx of pathogens from point and non-point

sources

Accumulation of pathogens in

environmental reservoirs

Removal of pathogens e.g. from die-off,

infiltration

Transport of pathogens in overland flow

Mobilization of pathogens (raindrop or

runoff erosion)

637

638

639

640

641

642

643

644

645

646

647

648

649

650

651

652

653

654

655

656

657

658

Page 42: opus.bath.ac.ukopus.bath.ac.uk/53573/1/Collender_disease_risks_inla… · Web viewopus.bath.ac.uk

yield from individual storm events; MUSLE incorporates antecedent soil moisture and erosion by runoff

into its estimations (eqn 24) (Renard, et al. 1991, Williams, et al. 2008):

E=α (V r o24∗Qmax∗A ls)β∗cserd∗cLS∗ccrm∗ccp (24 )

where V r o24 is the daily flow volume; Qmax is the maximum rate of runoff/discharge; Als is the land

surface area from which sediments are being transported; and α and β are calibrated terms to adapt the

equation to local conditions.

Of the models reviewed here for pathogen transport in runoff, COLI and SWAT rely on the MUSLE for

simulation of sediment transport and associated microbiological transport (Neitsch, et al. 2011, Walker, et

al. 1990). The implementations of WATFLOOD by Dorner et al. (2006) and Wu et al. (2009) directly

calculate pathogen mobilization and transport by rainfall impact and overland flow using the Hartley

model, an event-oriented expression that models rainfall-driven erosion using concepts from the USLE

and employs a stream power function to estimate erosion by overland flow (Hartley 1987):

Pro=Prf +Pof (25 )

Prf=R rf∗(11.9+8.7∗log ( R rf ))∗(1−c gc )∗ccan∗c perd (26 )

Pof =

60c ff

∗ρ∗Q

2∗sls

(27 )

where Prois the total number of pathogens transported in runoff; Prf is the number of pathogens

mobilized by raindrop impact; Pof is the number of pathogens mobilized and transported by overland

flow; Rrf is the rate or intensity of rainfall; c gc is a factor accounting for the effect of vegetative ground

42

659

660

661

662

663

664

665

666

667

668

669

670

671

672

673

674

675

676

677

Page 43: opus.bath.ac.ukopus.bath.ac.uk/53573/1/Collender_disease_risks_inla… · Web viewopus.bath.ac.uk

cover; ccan is a factor accounting for the effect of the canopy layer; c perd is a pathogen erodibility factor;

c ff is an empirical friction coefficient; ρ is the density of water; Q is the overland flow discharge; and sls

is the overland slope.

The treatment of sediment-associated pathogen mobilization in HSPF is based on a physically motivated

conceptual erosion model originating with the work of Negev (1967) and incorporating aspects of models

proposed by Meyer and Wischmeier (1969) and Onstad and Foster (1975) (eqns 28-32):

Pro=C pat h sedls

∗SE Dro=Cpat hse dls

∗( SE Dof +SE Drf ) (28 )

SE D of=αscour∗Qβ scour (29 )

{SE Drf =SE Dd etrfSE Dd etrf

<C se dmax

SE Drf =C sedmaxSE D detrf

≥ C sed max

(30 )

SE Dd et rf=(1−pc tcover )∗cman∗α sdet∗Rrf

β sdet (31 )

C sed max=α cap∗Qβcap (32 )

where C pat hsedls

is the concentration of pathogens in sediments on the land surface; SE D ro is the mass of

sediment transported in runoff; SE Dof is the mass of sediment eroded and transported by overland flow;

SE D rf is the mass of sediment mobilized by rainfall and transported in overland flow; α scour and βscour

are calibrated factors representing the contribution of scour by overland flow to sediment transport (α sdet,

43

678

679

680

681

682

683

684

685

686

687

688

689

690

Page 44: opus.bath.ac.ukopus.bath.ac.uk/53573/1/Collender_disease_risks_inla… · Web viewopus.bath.ac.uk

βsdet, α cap and βcap are analogous factors for sediment detachment by rainfall and transport capacity of

runoff, respectively); pc t cover represents the fraction of land protected from raindrop impacts by snow,

vegetative, or other cover; and cman is a factor representing land management practices.

Expressions for pathogen transport that do not invoke sediment mobilization are present in the

implementation of IHACRES by Ferguson et al. (2007), which specifies pathogen mobilization as a

function of excess rainfall, and the implementation of WAMVIEW by Tian et al. (2002), which expresses

the proportion of deposited pathogens transported to streams as a function of runoff volume, overland

flow distance, and calibrated coefficients. SWAT also permits simulation of pathogen transport without

consideration of sediment mobilization, using a simple function of runoff and a user-defined partitioning

coefficient (Neitsch, et al. 2011). Meanwhile, HSPF and the implementation of IHACRES by Ferguson et

al. express pathogen transport independent of sediments following assumptions of first-order kinetics

dictated by runoff and effective rainfall, respectively (eqn 33) (Ferguson, et al. 2007, Bicknell, et al.

1996):

dPdt

=−Q∗cwsh∗P ,∫0

∆t dPdt

=P∗(1−e−Q∗cwsh ) (33 )

where cwsh represents the susceptibility of the pathogen to washoff, and all other parameters are as

defined above.

Many physical processes are implicit in conceptual and empirical rainfall-runoff model parameters,

optimized values of these parameters often result in acceptable levels of agreement between modelled and

observed discharges or concentrations of indicator organisms. Thus, it is difficult to provide an

overarching assessment of the ability of such models to represent runoff conditions unique to flood

44

691

692

693

694

695

696

697

698

699

700

701

702

703

704

705

706

707

708

709

710

711

Page 45: opus.bath.ac.ukopus.bath.ac.uk/53573/1/Collender_disease_risks_inla… · Web viewopus.bath.ac.uk

events. In practice, models that rely on calibration of extensive sets of parameters are generally flexible

enough to be adapted to various watersheds and meteorological conditions; however, the data

requirements to do so may be costly or prohibitive, as has been noted for implementations of HSPF and

SWAT (Borah and Bera 2003, Gassman, et al. 2014). Failure to adequately calibrate models such as the

MUSLE, SWAT and HSPF, which include a large number of crucial user-defined parameters, often

results in large errors and very poor predictions of flow and transport (Borah and Bera 2003, Sadeghi, et

al. 2014), especially during periods of high flow (Borah and Bera 2003). Even then, issues of parameter

non-identifiability may prevent such high-dimensional models from generating good predictions outside

of the calibration range, and from providing useful information on the hydrological processes underlying

an area’s response to rainfall (Beven 2006, Freni, et al. 2009). Transport models based off the MUSLE

have been found to perform best when parameterized with directly measured runoff and peak flow data,

raising doubts about the reliability of models which apply the MUSLE using internally calculated

estimates of runoff and streamflow, such as SWAT (Sadeghi, et al. 2014). Properties of models for

pathogen transport in runoff are summarized in Table 8.

Table 8: Rainfall–Runoff Transport Model Characteristics

ModelMethods for

calculating excess rainfall

Sediment transport Pathogen transport

Watershed applications References

COLI SCS curve number MUSLE Sediment-associated transport

Temperate watershed,1153 km2

(Walker, et al. 1990)

IHACRES/ FERGUSON

ET AL.

Empirical non-linear loss module NA

First order kinetic function of

effective rainfall

Temperate watershed,

~90000 km2

(Ferguson, et al. 2007)

WATFLOOD/ DORNER/ WU

Process-based, infiltration excess model Hartley model Sediment-

analogous transport

Temperate, rural

watersheds,130-187 km2

(Dorner, et al. 2006, Wu, et al. 2009)

SWAT

Process-based model, SCS curve number, infiltration excess

options

MUSLESediment-

associated or direct transport

Many applications (Neitsch, et al. 2011)

HSPF Process-based, infiltration excess model

Power laws for rainfall or runoff-

based erosion, limited by supply or transport capacity

Sediment-associated transport or first order kinetic function of runoff

Many applications (Bicknell, et al. 1996)

STWIR Process-based, infiltration excess model

Rainfall splash erosion according to

formula of Meyer

Advection-diffusion equation

Many applications

(Guber, et al. 2012, Meyer and

Wischmeier 1969,

45

712

713

714

715

716

717

718

719

720

721

722

723

724

725

726

Page 46: opus.bath.ac.ukopus.bath.ac.uk/53573/1/Collender_disease_risks_inla… · Web viewopus.bath.ac.uk

and Wischmeier

Several shear stress-dependent physically

based models for erosion by runoff

Goodrich, et al. 2010)

WAMVIEW/ TIAN

Options include modified SCS curve number, Saturation

excess model, or special cases algorithm with

user-defined parameters

Delivery ratio method incorporating landscape factors and

runoff

Delivery ratio method based on

runoff, distance to stream, and user-

defined coefficients

Temperate agricultural watersheds

0.49-0.95 km2

(Tian, et al. 2002, Soil & Water Engineering

Technology Inc.)

WEPP Process-based infiltration excess model

Shear-stress dependent

physically-based erosion from runoff

Advection equation with stochastic

partition between mobile and

immobile states

Developed for use on fields ~2.6-8.1 km2

(Yeghiazarian, et al. 2006, Bhattarai, et al. 2011, Flanagan and

Nearing 1995)

46

727

Page 47: opus.bath.ac.ukopus.bath.ac.uk/53573/1/Collender_disease_risks_inla… · Web viewopus.bath.ac.uk

5. Discussion

5.1 Gaps in pathogen fate and transport modeling practice relevant to flood conditions

Several phenomena relevant to pathogen fate and transport during floods have not been addressed in

current models. These include process descriptions of the mobilization of pathogens from sanitation

facilities in inundated areas, the role of elevated levels of dissolved organic matter in altering pathogen

transport and persistence, and the eventual depletion of pathogen sources that may occur during sustained

or consecutive flood events. Representation of these processes, especially the latter two, within integrated

models may entail greater model complexity than currently available data are capable of supporting.

Nonetheless, mathematical descriptions of these phenomena are a necessary prerequisite to their eventual

incorporation into model frameworks as deterministic processes or sources of uncertainty.

Mobilization of pathogens from inundated facilities, such as pit latrines, may be considered as either a

turbulent diffusion/resuspension process, in which flow over a deep compartment results in passive or

active movement of pathogens from the compartment into the overlying waters, or as the result of

structural damage to the facility caused by sufficiently deep and swift flow, resulting in a rapid release of

pathogens from within the facility. Controlled experiments to measure the release of indicator organisms

from replica latrines in artificial compound channels may provide a suitable baseline for efforts to model

these phenomena.

With respect to the effects of heightened concentrations of dissolved organic matter on pathogen fate and

transport, current knowledge indicates that a decrease in pathogen retention in soils and sediments is

likely, while persistence of bacterial and viral pathogens may be enhanced, or, through the action of

antagonistic microbial communities, diminished (Bradford, et al. 2013). Literature on bacterial die-off

indicates that high nutrient concentrations are associated with extended periods of maintenance or growth

prior to population decline. This maintenance period is not captured by first-order exponential decay

47

728

729

730

731

732

733

734

735

736

737

738

739

740

741

742

743

744

745

746

747

748

749

750

751

752

Page 48: opus.bath.ac.ukopus.bath.ac.uk/53573/1/Collender_disease_risks_inla… · Web viewopus.bath.ac.uk

equations used in most transport models, although a piecewise model has been proposed to incorporate it

(eqn 31; Darakas 2002):

C path (t )={ C path(0) for t ≤ t d

C path (0 )∗e−Kd∗(t−td ) for t >t d

(31 )

where C path ( t ) is the concentration of organisms at time t ; t d is the duration of the maintenance phase

prior to die-off; Kd is the die-off rate constant. Dissolved Organic Carbon has been suggested as a

suitable indicator for dissolved nutrients (Hipsey, et al. 2008). Where DOC data and, ideally, estimates of

total microbial metabolism (e.g. the biological oxygen demand (BOD)), are available, it may be possible

to represent the maintenance phase as a function of these quantities. Pathogen die-off experiments in

which DOC and BOD are measured and evaluated as determinants of the length of the maintenance phase

will be a prerequisite for its incorporation into fate and transport models.

When flooding or extreme precipitation is sufficiently severe, or occurs frequently, the amount of

pathogens available for transport from the land surface may eventually be exhausted (Muirhead, et al.

2004, Nagels, et al. 2002). While only a few models, such as SWAT (Neitsch, et al. 2011), HSPF

(Bicknell, et al. 1996), and the Dorner et al. model (2006), have incorporated simulation of supply-limited

pathogen transport to date, it is conceptually simple to incorporate rates of pathogen shedding and

persistence in relevant environmental compartments into any rainfall-runoff or land-surface model,

although a major data gathering effort would be necessary to support any such implementation, and

considerable uncertainty regarding appropriate parameter values and their variability would almost

certainly remain. This approach would provide a basis to estimate the total population of pathogens

available for transport from a compartment at any time step. The depletion of pathogens available for

transport from each compartment in response to consecutive flooding events would allow models to

48

753

754

755

756

757

758

759

760

761

762

763

764

765

766

767

768

769

770

771

772

773

774

Page 49: opus.bath.ac.ukopus.bath.ac.uk/53573/1/Collender_disease_risks_inla… · Web viewopus.bath.ac.uk

represent the flushing of fecal contamination by heavy runoff and high flows, which has been advanced as

an important mechanism modulating the effects of precipitation on disease transmission (Carlton, et al.

2014).

Finally, it is worth mentioning that pathogens, which have largely been discussed as an abstract,

monolithic group of suspended constituents in this review, differ with regards to size, density, resilience

to various environmental challenges, and particle affinity. The vast majority of microbial transport

modeling has been conducted using fecal indicator bacteria, which are unlikely to behave similarly to

protozoans and viruses (Ashbolt, et al. 2001). Even within a pathogen class, such as bacteria, there is

considerable heterogeneity with respect to the above mentioned characteristics (Jenkins, et al. 2011,

Schillinger and Gannon 1985). As more robust microbiological fate and transport models emerge, they

should be parameterized and validated using pathogen-specific data, starting with indicator organisms

known to be well-correlated with transport characteristics of important viral, bacterial, and protozoan

pathogens (Ferguson, et al. 2003, Ashbolt, et al. 2001).

49

775

776

777

778

779

780

781

782

783

784

785

786

787

788

789

Page 50: opus.bath.ac.ukopus.bath.ac.uk/53573/1/Collender_disease_risks_inla… · Web viewopus.bath.ac.uk

5.2 Where and when should models incorporate increased process complexity?

Models discussed in this review span a range of complexity in process representation, from the almost

entirely empirical, such as IHACRES, to spatially distributed, physically based models, such as

TELEMAC 2D. While increasing model complexity allows for a more comprehensive description of

hydrological and transport processes, it must be acknowledged that the extra parameters entailed by

complex process representation and fine spatial discretization will often result in parameter non-

identifiability and model equifinality, stymying the initial intent of a more detailed process representation,

as there is no guarantee that the ‘optimal’ parameter values of a calibrated model truly represent the

processes driving hydrology and pathogen transport (Beven 2006). This situation is further exacerbated

by issues of data quantity and quality, especially in the developing world, where vulnerability to flood-

related diseases is greatest. Thus, model parsimony is as much a concern as detailed process

representation, and detailed analysis of structural and statistical uncertainties is critical in model

applications.

50

790

791

792

793

794

795

796

797

798

799

800

801

802

803

Page 51: opus.bath.ac.ukopus.bath.ac.uk/53573/1/Collender_disease_risks_inla… · Web viewopus.bath.ac.uk

The relative effects of parameters and process representations in fate and transport modeling are not well

understood, and are likely to be specific to the pathogen of interest, geographical location, flood event

magnitude, and perhaps even the flow domain (e.g. different relative effects for the channel and the

floodplain) (Sommerfreund, et al. 2010). Some model capabilities, such as mass-and-momentum

conservative exchanges of flow between channels and floodplains, adequate accounting for pathogen

sources in upstream watersheds, and some capacity to represent dispersion and/or retention and release of

transported pathogens, seem likely to improve estimates of pathogen transport associated with floods,

assuming adequate data on channel and floodplain bathymetry/topography and influxes to the flow

domain are available. However, the degree to which more or less detailed process representations of

pathogen suspension, retention, sedimentation, dispersion within flows, and changes in pathogen

persistence due to the effects of suspended solids and dissolved organic matter will affect model skill in

estimating pathogen transport is likely to vary by model application. In order to arrive at parsimonious

model parameterizations, sensitivity analysis techniques can be used to selectively remove parameters

with little impact on model outputs from calibration or randomization processes, or to reduce the

dimensionality of models through principle components analysis and similar techniques (see e.g. Freni et

al. (2009), Sommerfreund et al. (2010)). Alternative model structures and overall uncertainty can be

assessed within Bayesian frameworks, such as the General Likelihood Uncertainty Estimation (GLUE)

method of Beven et al. (Beven 2007, Beven and Binley 1992). To characterize the range of conditions or

research goals for which simplified process representations will be sufficient or preferable, comparative

studies are needed which examine model behaviors over a wide range of rainfall regimes, topography,

soil and land use characteristics, and pathogen source scenarios. While there is still much to be learned

about generalized hydrodynamics and transport during floods, such studies would also ideally utilize time

series of observed indicator organism concentrations as benchmarks for model performance.

51

804

805

806

807

808

809

810

811

812

813

814

815

816

817

818

819

820

821

822

823

824

825

826

827

Page 52: opus.bath.ac.ukopus.bath.ac.uk/53573/1/Collender_disease_risks_inla… · Web viewopus.bath.ac.uk

5.3 Summary of properties of reviewed fate and transport models relevant to flood conditions

Existing microbial fate and transport models vary widely in their approaches to simulating the

hydrodynamic flows which drive pathogen transport, with potential consequences for their validity under

flood conditions. Hydrodynamic transport models cover a spectrum of physical complexity, from 1D

steady/uniform flow models (HSPF, WATFLOOD, SWAT, QUAL2K, WAMVIEW, QUAL2E), to 1D

models using the dynamic wave equation (or close approximations of it) to characterize unsteady, non-

uniform flow (SOBEK, Kazama et. al (2012)), to 2D or higher models (HEMAT, TELEMAC, DIVAST,

EFDC) capable of representing secondary flows within channels and inundated areas. SOBEK,

WATFLOOD, SWAT, the Kazama model, and DUFLOW have the capacity to model mass-conservative

exchanges of flow (and therefore pathogens) between channels and floodplains. However, of the reviewed

models, only TELEMAC and HEMAT attempt conservation of momentum in channel-floodplain

exchanges. SWAT and WATFLOOD make simplifying assumptions regarding floodplain topography

which limits their realism in modeling overbank flows.

The parameterization of source, sink, and dispersion terms included in hydrodynamic transport equations

(e.g., eqn 9) also varies widely across existing microbial transport models. Pathogen mobilization from

bed sediments is not accounted for in SOBEK, QUAL2K, FASTER, DUFLOW, TELEMAC 2D, SLIM,

QUAL2E, or HEMAT. Simple models for the mobilization of sediment-associated pathogens are present

in the implementations of WATFLOOD by Dorner et al. (2006) and Wu et al. (2009), while

erosion/deposition models based on concepts of sediment transport capacity are used in WAMVIEW, and

in SWAT’s default configuration. Sediment erosion and deposition is modeled with a higher degree of

physical process representation, based on shear stresses in the model of Yakirevich et al. (2013), as well

as in DIVAST, HSPF, SWAT’s optional configurations, EFDC, and MOBED. Removal of pathogens

from the water column by sedimentation is not accounted for in SOBEK, the Kazama model, FASTER,

DUFLOW, TELEMAC, QUAL2E, or HEMAT. Sedimentation is modeled based on constant pathogen

settling velocity in WATFLOOD, QUAL2K, and SLIM. Variable dispersion parameters are neglected in

52

828

829

830

831

832

833

834

835

836

837

838

839

840

841

842

843

844

845

846

847

848

849

850

851

852

853

Page 53: opus.bath.ac.ukopus.bath.ac.uk/53573/1/Collender_disease_risks_inla… · Web viewopus.bath.ac.uk

SOBEK, HSPF, WATFLOOD, WAMVIEW, DUFLOW, and MOBED. Simpler formulations relating

dispersion to flow parameters are present in the Yakirevich et al. (2013) model, SLIM, and QUAL2E,

while more complex empirical models are applied in DIVAST, QUAL2K, FASTER, and full turbulence

modeling is available in TELEMAC and EFDC. Regarding the mechanisms through which flood

conditions are likely to affect microbial persistence, only the Kazama model includes parameters to model

the effect of suspended solids on pathogen survival.

The validity of models for the transport of pathogens in runoff under flood conditions is difficult to assess

from a theoretical standpoint. Many rainfall-runoff models, including the implementation of IHACRES

by Ferguson et al. (2007), WATFLOOD by Dorner et al. (2006) and Wu et al. (2009), COLI, GIBSI,

SWAT, HSPF, and WAMVIEW, rely on empirical or conceptual functions to relate pathogen loading of

receiving streams to total runoff or effective rainfall. As a result, the accuracy of estimates for pathogen

transport in runoff will depend on the spatial and temporal resolutions at which these models are run, and

the selection of appropriate parameters for the regional and meteorological conditions for which they are

implemented. Spatially distributed rainfall-runoff transport models, such as WEPP and STWIR, explicitly

account for transient exchanges of pathogens between overland flow, the land surface, and the subsurface

via calibrated rate parameters incorporated into hydrodynamic equations. However, these equations are

designed for implementation in relatively small areas and require extensive collection of data on

catchment characteristics in order to produce accurate results. Properties of the reviewed models are

briefly summarized in Table 9.

Table 9: Summary of capabilities of reviewed models

ModelLand surface simulation

Streamflow simulation

Floodplain simulation

Channel-floodplain exchange

DownstreamPathogen routing

Pathogen settling and resuspension

Pathogen dispersion

Persistencefactors

HSPF Semi-distributed

1D storage routing NA NA Mixing cell

Simplified shear stress formulation

NA

Temperature and radiation, particle association, deposition

WATFLOOD/ DORNER/ WU

Semi-distributed

1D storage routing

1D storage routing (simplified geometry)

Mass conserved

Mixing cell Power law suspension based on discharge,

NA Season

53

854

855

856

857

858

859

860

861

862

863

864

865

866

867

868

869

870

871

872

873

874

Page 54: opus.bath.ac.ukopus.bath.ac.uk/53573/1/Collender_disease_risks_inla… · Web viewopus.bath.ac.uk

constant deposition at particle settling velocity

IHACRES/ FERGUSON Lumped NA NA NA NA NA NA Pathogen type

SWAT Semi-distributed

1D storage or Muskingum routing

1D storage or Muskingum routing (simplified geometry)

Mass conserved NA

Sediment carrying capacity formulation, various degress of physical rigor available

NA

Temperature, pathogen type, environmental compartment

KINEROS2 /STWIR

Fully distributed hydrodynamic

1D kinematic wave

1D kinematic wave

Mass conserved

Advection-diffusion equation

NA

Time-invariant calibrated parameter

Deposition, environmental compartment

COLI Lumped NA NA NA NA NA NA Temperature

WAMVIEWFully distributed conceptual

1D storage routing NA NA Mixing cell

Simple function of daily flow volume

NA Temperature, radiation

WEPP

Fully distributed hydrodynamic

NA NA NA NA NA NA None

DIVAST NA2D dynamic wave

2D dynamic wave

Mass and momentum conserved

Advection-diffusion equation

Shear stress formulation

Calculated from depth, flow velocities, and roughness coefficient

None

SOBEK/D-WATER QUALITY

NA1D dynamic wave

2D dynamic wave

Mass conserved

Advection-diffusion equation

NA NASalinity, temperature, radiation

KAZAMA ET AL. NA

1D dynamic wave

2D dynamic wave without convective acceleration

Mass conserved

Advection equation NA NA

Radiation, water depth, suspended solids

Yakirevich et al. NA

1D dynamic wave

NA NAAdvection-diffusion equation

Shear stress formulation

Calculated from flow velocity

Deposition

QUAL2K NA 1D storage routing NA NA Mixing cell

No suspension, deposition dependent on depth and particle settling velocity

Calculated from water depth, channel slope, channel width, and flow velocity

Temperature, radiation, water depth

FASTER NA 1D dynamic NA NA Advection-

diffusion NA Calculated from

Day and night, coastal or fresh

54

Page 55: opus.bath.ac.ukopus.bath.ac.uk/53573/1/Collender_disease_risks_inla… · Web viewopus.bath.ac.uk

wave equation

water depth, channel width, flow velocity, shear velocity

waters, dry or wet weather

EFDC NA1, 2, or 3D dynamic wave

NA NAAdvection-diffusion equation

Shear stress formulation

Explicit turbulence modeling

Temperature, radiation, salinity, pH, particle association, deposition

DUFLOW NA1D dynamic wave

1D dynamic wave

Mass conserved

Advection-diffusion equation

NA NA None

TELEMAC NA2 or 3D dynamic wave

2 or 3D dynamic wave

Mass and momentum conserved

Advection-diffusion equation

NA

Explicit turbulence modeling

None

SLIM NA1D dynamic wave

NA NAAdvection-diffusion equation

No suspension, deposition dependent on depth, particle settling velocity, and pathogen concentration

NA Temperature

MOBED NA1D dynamic wave

NA NA Mixing cellSimplified shear stress formulation

NA NA

QUAL2E-GIBSI

Semi-distributed

1D storage routing NA NA

Advection-diffusion equation

NA

Calculated from roughness, mean flow velocity and depth of flow

NA

HEMAT NA2D dynamic wave

2D dynamic wave

Mass and momentum conserved

Advection-diffusion equation

NA NA Day and night

55

875

Page 56: opus.bath.ac.ukopus.bath.ac.uk/53573/1/Collender_disease_risks_inla… · Web viewopus.bath.ac.uk

5.4 Improving data availability and integrating modeling and field efforts for vulnerable regions

Parameterizing pathogen fate and transport models, especially those with more complex representation of

transport processes, requires extensive environmental data (Aksoy and Kavvas 2005). Unfortunately,

many areas at heightened risk of flood-related infectious disease transmission are resource poor, and

unlikely to have extensive high-quality environmental datasets, and in situ collection of data during flood

events may be impossible. As a result, researchers may seek to produce adequate models for vulnerable

areas without extensive in situ data on soils, channel morphology, and/or streamflow records, not to

mention the regular measurements of pathogen or sediment concentrations needed to calibrate fate and

transport models (Coffey, et al. 2010). Remotely sensed Synthetic Aperture Radar (SAR) and LIDAR

images of flood extent and topographical features have proven useful for parameterizing and validating a

number of flood inundation models at high spatial resolutions (Bates 2004, Horritt and Bates 2002, Bates,

et al. 2003, Straatsma and Baptist 2007, Smith 1997, Patro, et al. 2009, Bates, et al. 1997). However, the

use of such image-based techniques to describe the dynamics of inundation over the duration of a flood

event, and thus establish credible estimates of flow velocities and other transport-relevant parameters,

requires high resolution imagery, both spatially and temporally (Bates 2012, Di Baldassarre and

Uhlenbrook 2012). In addition to inundation extent, it may also be possible to identify proxies for

pathogen concentrations and transport that can be measured remotely, such as turbidity (Bradford and

Schijven 2002, Schijven, et al. 2004). Researchers have also proposed methods for imputing channel

geometry and bathymetry, crucial parameters for hydrodynamic models that are often difficult to obtain in

developing countries (Wood, et al. 2016). Given pervasive limitations on obtaining in situ data, especially

in developing country settings, further advances are needed in methods for integrating remote sensing

data and models to impute missing environmental data. These activities should be accompanied by firm

advocacy for the expansion of reliable ground-based environmental monitoring stations in vulnerable

global settings.

56

876

877

878

879

880

881

882

883

884

885

886

887

888

889

890

891

892

893

894

895

896

897

898

899

900

Page 57: opus.bath.ac.ukopus.bath.ac.uk/53573/1/Collender_disease_risks_inla… · Web viewopus.bath.ac.uk

Data collected by environmental field studies can be an important source of information for those

developing models of flood-related infection risks, particularly in areas where standardized, routine

monitoring is limited or absent. There is often poor linkage, however, between parameter and validation

data needed by modelers and the data typically collected in environmental field studies. To a certain

extent, a disconnect is unavoidable, since many model parameters and outputs are ‘effective’ parameters

representing some averaging of characteristics over a spatiotemporal discretization, whereas field

measurements are generally taken as point measurements (Beven 2010b). Other issues arise because the

design and corresponding data collection protocols for field studies are motivated by highly specific

research questions, and the intensity and scope of data collection are often limited by sparse sampling

equipment and constrained study logistics. Yet both field-based researchers and the modeling community

stand to benefit from greater coordination of field studies with modeling efforts. Focusing in situ data

collection on key parameters that underlie hydrological and microbial transport models would be of great

benefit for advancing model capabilities, while also providing information useful for the design of future

field experiments. Moreover, strengthening collaborations between modelers and field researchers can

further the development of models capable of providing mechanistic validation of associations discovered

in the field, and can thus increase the interplay between theory, observation and experimentation.

5.5 Perspective for risk-based engineering design

The application of hydrological and hydrodynamic pathogen fate and transport models to risk assessment

may be viewed as an example of a decision-making processes involving engineering design. In this sense,

fate and transport estimation is analogous to the design of hydraulic structures such as flood defenses and

drainage systems, in that the frequency and severity of future events is inherently uncertain. Design flood

events used in hydraulic engineering are derived either from direct statistical analysis of long-term

records (typically in excess of 20 years) or by considering the joint probability of several factors involved

in flood generation such as rainfall and antecedent soil moisture conditions coupled with a deterministic

watershed model (Svensson, et al. 2013). As long-term observations of pathogens are rare and unlikely to

57

901

902

903

904

905

906

907

908

909

910

911

912

913

914

915

916

917

918

919

920

921

922

923

924

925

926

Page 58: opus.bath.ac.ukopus.bath.ac.uk/53573/1/Collender_disease_risks_inla… · Web viewopus.bath.ac.uk

be available for most practical applications, particularly in developing countries, it seems that the best

way forward for providing risk-based estimates of pathogen concentrations is to include the deterministic

and random components known to control pathogen concentrations into the joint probability approach.

However, as discussed in Section 4.7, there is still insufficient scientific evidence available to reliably

link, for example, pathogen mobilization to rainfall intensity and watershed characteristics.

Consequently, a quantitative tool for providing risk-based estimates of pathogen concentrations must still

be considered a future ambition.

58

927

928

929

930

931

932

933

934

Page 59: opus.bath.ac.ukopus.bath.ac.uk/53573/1/Collender_disease_risks_inla… · Web viewopus.bath.ac.uk

6. Conclusion

Given the computational expense, often severe mismatches between data requirements and data

availability, and at least partially irreducible uncertainties associated with microbial fate and transport

modeling in the context of flood events, one might well ask what is to be gained in the endeavor. It is

clear that, once implemented, flood and microbial fate and transport models should not be taken as

deterministic predictions of risk, especially in the contexts of rare and extreme events and hydrological

non-stationarity due to climate change. Despite the aforementioned difficulties, fate and transport models,

as platforms for knowledge integration and synthesis, have the potential to provide much-needed

information to guide research and risk mitigation efforts. In this context, flood and microbial transport

modeling should be viewed as an iterative project, where initial applications may help distinguish

between processes to which model outputs are more or less sensitive in various contexts, as well as

geographical areas estimated to be at high or low risk regardless of model specification, and others for

which uncertain or poor predictions may motivate directed data collection or modifications to existing

theory or model structures (Sommerfreund, et al. 2010).

59

935

936

937

938

939

940

941

942

943

944

945

946

947

948

949

Page 60: opus.bath.ac.ukopus.bath.ac.uk/53573/1/Collender_disease_risks_inla… · Web viewopus.bath.ac.uk

Initially, advancing models of microbial risks associated with flood events will require addressing several

areas of incomplete knowledge. New modeling techniques should be developed and validated for

processes that have not previously received attention in the hydrological modeling literature, including

the mobilization of pathogens from flooded sanitation facilities. In order to reduce sizable uncertainties

due to existing data limitations, especially for developing countries, it will be necessary to support and

incorporate multiple sources of information, including continuing to pair and enhance existing models

with more reliable and comprehensive field measurements, as well as new and more extensive remotely

sensed data, and innovative measurement techniques, especially for measuring the properties of overbank

flows. Finally, incorporation of probabilistic model structures (e.g. hierarchical models to specify

distributed roughness coefficients from a common distribution (Rode, et al. 2010)), parameter

identifiability analysis, and ensemble modeling techniques, such as GLUE, will be critical, allowing

researchers to incorporate and refine prior knowledge, distinguish irreducible and reducible uncertainties

in transport parameters, and test various models as alternative hypotheses.

7. Acknowledgements

This work was supported in part by the Chemical, Bioengineering, Environmental, and Transport Systems

Division of the National Science Foundation under grant no. 1249250, by the Division of Earth Sciences

of the National Science Foundation under grant nos. 1360330 and 1646708, by the National Institute for

Allergy and Infectious Disease (K01AI091864) and by the National Institutes of Health/National Science

Foundation Ecology of Infectious Disease program funded by the Fogarty International Center (grant

R01TW010286). Financial support for OC’s contributions to this publication come from the UK

Engineering and Physical Sciences Research Council’s (EPSRC) Water Informatics Science &

Engineering (WISE) Centre for Doctoral Training program (Grant Reference EP/L016214/1) and the

University of Bath International Mobility fund.

60

950

951

952

953

954

955

956

957

958

959

960

961

962

963

964

965

966

967

968

969

970

971

972

973

974

Page 61: opus.bath.ac.ukopus.bath.ac.uk/53573/1/Collender_disease_risks_inla… · Web viewopus.bath.ac.uk

The authors would like to thank Wen Yang, Jedidiah Snyder, Marc Stieglitz, and Kevin Zhu for their

input on early drafts of this manuscript.

61

975

976

977

Page 62: opus.bath.ac.ukopus.bath.ac.uk/53573/1/Collender_disease_risks_inla… · Web viewopus.bath.ac.uk

References:

EM-DAT (2014). The OFDA/CRED international disaster database.Jonkman, S.N. (2005). Global perspectives on loss of human life caused by floods. Natural Hazards 34, 151-175.United Nations (2012). World urbanization prospects: the 2011 revision. New York.IPCC (2013). Working group I contribution to the IPCC fifth assessment report. Climate Change 2013: The physical science basis. Cambridge, United Kingdom and New York, NY, USA.Ahern, M., Kovats, R.S., Wilkinson, P., Few, R. and Matthies, F. (2005). Global health impacts of floods: epidemiologic evidence. Epidemiolgic Reviews 27, 36-46.Jonkman, S.N., Vrijling, J.K. and Vrouwenvelder, A.C.W.M. (2008). Methods for the estimation of loss of life due to floods: a literature review and a proposal for a new method. Natural Hazards 46, 353-389.Hunter, P.R. (2003). Climate change and waterborne and vector-borne disease. Journal of Applied Microbiology 94, 37S-46S.Alderman, K., Turner, L.R. and Tong, S. (2012). Floods and human health: a systematic review. Environment International 47, 37-47.McBride, G.B. and Mittinty, M.N. (2007). Explaining differential timing of peaks of a pathogen versus a faecal indicator during flood events. International Congress on Modelling and Simulation (MODSIM07). Christchurch, New Zealand.Walters, E., Schwarzwälder, K., Rutschmann, P., Müller, E. and Horn, H. (2014). Influence of resuspension on the fate of fecal indicator bacteria in large-scale flumes mimicking an oligotrophic river. Water Research 48, 466-477.Singh, V.P. and Woolhiser, D.A. (2002). Mathematical modeling of watershed hydrology. Journal of Hydrologic Engineering 7, 270-292.Miller, J.D., Kim, H., Kjeldsen, T.R., Packman, J., Grebby, S. and Dearden, R. (2014). Assessing the impact of urbanization on storm runoff in a peri-urban catchment using historical change in impervious cover. Journal of Hydrology 515, 59-70.Singh, V.P. (1996). Kinematic wave modeling in water resources: surface-water hydrology. New York, N.Y.: John Wiley and Sons.Klemeš, V. (1986). Operational testing of hydrological simulation models. Hydrological Sciences Journal 31, 13-24.Beven, K. (2006). A manifesto for the equifinality thesis. Journal of Hydrology 320, 18-36.Beven, K. (2001). How far can we go in distributed hydrological modelling? Hydrol. Earth Syst. Sci. 5, 1-12.Jamieson, R., Joy, D.M., Lee, H., Kostaschuk, R. and Gordon, R. (2005). Transport and deposition of sediment-associated Escherichia coli in natural streams. Water Research 39, 2665-2675.de Brauwere, A., Ouattara, N.K. and Servais, P. (2014). Modeling fecal indicator bacteria concentrations in natural surface waters: a review. Critical Reviews in Environmental Science and Technology 44, 2380-2453.Sommerfreund, J., Arhonditsis, G.B., Diamond, M.L., Frignani, M., Capodaglio, G., Gerino, M., Bellucci, L., Giuliani, S. and Mugnai, C. (2010). Examination of the uncertainty in contaminant

62

978

979980981982983984985986987988989990991992993994995996997998999

10001001100210031004100510061007100810091010101110121013101410151016101710181019

Page 63: opus.bath.ac.ukopus.bath.ac.uk/53573/1/Collender_disease_risks_inla… · Web viewopus.bath.ac.uk

fate and transport modeling: A case study in the Venice Lagoon. Ecotoxicology and Environmental Safety 73, 231-239.Neal, J.C., Bates, P.D., Fewtrell, T.J., Hunter, N.M., Wilson, M.D. and Horritt, M.S. (2009). Distributed whole city water level measurements from the Carlisle 2005 urban flood event and comparison with hydraulic model simulations. Journal of Hydrology 368, 42-55.Ghimire, B. and Deng, Z. (2013). Hydrograph-based approach to modeling bacterial fate and transport in rivers. Water Research 47, 1329-1343.Wu, J., Rees, P., Storrer, S., Alderisio, K. and Dorner, S. (2009). Fate and transport modeling of potential pathogens: The contribution from sediments. Jounal of the American Water Resources Association 45, 35-44.Kazama, S., Aizawa, T., Watanabe, T., Ranjan, P., Gunawardhana, L. and Amano, A. (2012). A quantitative risk assessment of waterborne infectious disease in the inundation area of a tropical monsoon region. Sustainability Science 7, 45-54.Kazama, S., Hagiwara, T., Ranjan, P. and Sawamoto, M. (2007). Evaluation of groundwater resources in wide inundation areas of the Mekong River basin. Journal of Hydrology 340, 233-243.Bradford, S.A., Morales, V.L., Zhang, W., Harvey, R.W., Packman, A.I., Mohanram, A. and Welty, C. (2013). Transport and fate of microbial pathogens in agricultural settings. Critical Reviews in Environmental Science and Technology 43, 775-893.Costabile, P., Costanzo, C. and Macchione, F. (2013). A storm event watershed model for surface runoff based on 2D fully dynamic wave equations. Hydrological Processes 27, 554-569.Ikeda, S. and McEwan, I.K. (2009). Flow and sediment transport in compound channels: the experience of Japanese and UK research: Taylor & Francis.Nittrouer, J.A., Shaw, J., Lamb, M.P. and Mohrig, D. (2011). Spatial and temporal trends for water-flow velocity and bed-material sediment transport in the lower Mississippi River. Geological Society of America Bulletin.Schulz, M., Büttner, O., Baborowski, M., Böhme, M., Matthies, M. and von Tümpling, W. (2009). A dynamic model to simulate arsenic, lead, and mercury contamination in the terrestrial environment during extreme floods of rivers. CLEAN – Soil, Air, Water 37, 209-217.Elder, J.W. (1959). The dispersion of marked fluid in turbulent shear flow. Journal of Fluid Mechanics 5, 544-560.Fischer, H.B. (1975). Simple method for predicting dispersion in streams. Journal of the Environmental Engineering Division-ASCE 101, 453-455.Kashefipour, S.M. and Falconer, R.A. (2002). Longitudinal dispersion coefficients in natural channels. Water Research 36, 1596-1608.Bormann, H., Breuer, L., Giertz, S., Huisman, J. and Viney, N. (2009). Spatially explicit versus lumped models in catchment hydrology – experiences from two case studies. In P. Baveye et al. (eds.) Uncertainties in environmental modelling and consequences for policy making (NATO science for peace and security series c: Environmental security, 3-26): Springer Netherlands.Beven, K. (2010a). Environmental modelling: An uncertain future?: Taylor & Francis.Tayefi, V., Lane, S.N., Hardy, R.J. and Yu, D. (2007). A comparison of one- and two-dimensional approaches to modelling flood inundation over complex upland floodplains. Hydrological Processes 21, 3190-3202.

63

1020102110221023102410251026102710281029103010311032103310341035103610371038103910401041104210431044104510461047104810491050105110521053105410551056105710581059106010611062

Page 64: opus.bath.ac.ukopus.bath.ac.uk/53573/1/Collender_disease_risks_inla… · Web viewopus.bath.ac.uk

Pilotti, M., Maranzoni, A., Milanesi, L., Tomirotti, M. and Valerio, G. (2014). Dam-break modeling in alpine valleys. Journal of Mountain Science 11, 1429-1441.Bladé, E., Gómez-Valentín, M., Dolz, J., Aragón-Hernández, J.L., Corestein, G. and Sánchez-Juny, M. (2012). Integration of 1D and 2D finite volume schemes for computations of water flow in natural channels. Advances in Water Resources 42, 17-29.Borah, D.K., Arnold, J.G., Bera, M., Krug, E.C. and Liang, X.-Z. (2007). Storm event and continuous hydrologic modeling for comprehensive and efficient watershed simulations. Journal of Hydrologic Engineering 12, 605-616.Miller, J.E. (1984). Basic concepts of kinematic-wave models. Report No. 1302.Novak, P., Guinot, V., Jeffrey, A. and Reeve, D.E. (2010). Hydraulic modelling – an introduction: principles, methods and applications: Taylor & Francis.Kouwen, N. (2014). WATFLOODtm / WATROUTE hydrological model routing & flow forecasting system. Ontario, Canada.Martin, J.L. and McCutcheon, S.C. (1998). Hydrodynamics and transport for water quality modeling: Taylor & Francis.Koussis, A.D. (2009). Assessment and review of the hydraulics of storage flood routing 70 years after the presentation of the Muskingum method. Hydrological Sciences Journal 54, 43-61.Beven, K. (2007). Uncertainty in predictions of floods and hydraulic transport Transport phenomena in hydraulics, 5-20. Warsaw: Institute of Geophysics.Besio, G., Stocchino, A., Angiolani, S. and Brocchini, M. (2012). Transversal and longitudinal mixing in compound channels. Water Resources Research 48(12).Seckin, G., Mamak, M., Atabay, S. and Omran, M. (2009). Discharge estimation in compound channels with fixed and mobile bed. Sadhana 34, 923-945.Chaturongkasumrit, Y., Techaruvichit, P., Takahashi, H., Kimura, B. and Keeratipibul, S. (2013). Microbiological evaluation of water during the 2011 flood crisis in Thailand. Science of The Total Environment 463-464, 959-967.Funari, E., Manganelli, M. and Sinisi, L. (2012). Impact of climate change on waterborne diseases. Annali dell'Istituto Superiore di Sanità 48, 473-487.Bhavnani, D., Goldstick, J.E., Cevallos, W., Trueba, G. and Eisenberg, J.N.S. (2014). Impact of rainfall on diarrheal disease risk associated with unimproved water and sanitation. The American Journal of Tropical Medicine and Hygiene 90, 705-711.Baqir, M., Sobani, Z.A., Bhamani, A., Bham, N.S., Abid, S., Farook, J. and Beg, M.A. (2012). Infectious diseases in the aftermath of monsoon flooding in Pakistan. Asian Pac J Trop Biomed 2, 76-79.ten Veldhuis, J.A.E., Clemens, F.H.L.R., Sterk, G. and Berends, B.R. (2010). Microbial risks associated with exposure to pathogens in contaminated urban flood water. Water Research 44, 2910-2918.Massoud, M.A., Tarhini, A. and Nasr, J.A. (2009). Decentralized approaches to wastewater treatment and management: applicability in developing countries. Journal of Environmental Management 90, 652-659.Shimi, A.C., Parvin, G.A., Biswas, C. and Shaw, R. (2010). Impact and adaptation to flood: a focus on water supply, sanitation and health problems of rural community in Bangladesh. Disaster Prevention and Management: An International Journal 19, 298-313.

64

1063106410651066106710681069107010711072107310741075107610771078107910801081108210831084108510861087108810891090109110921093109410951096109710981099110011011102110311041105

Page 65: opus.bath.ac.ukopus.bath.ac.uk/53573/1/Collender_disease_risks_inla… · Web viewopus.bath.ac.uk

Benham, B.L., Baffaut, C., Zeckoski, R.W., Mankin, K.R., Pachepsky, Y.A., Sadeghi, A.M., Brannan, K.M., Soupir, M.L. and Habersack, M.J. (2006). Modeling bacteria fate and transport in watersheds to support TMDLs. Transactions of the ASABE 49, 987-1002.Blaustein, R.A., Pachepsky, Y.A., Shelton, D.R. and Hill, R.L. (2015). Release and removal of microorganisms from land-deposited animal waste and animal manures: a review of data and models. J Environ Qual 44, 1338-1354.Carlton, E.J., Eisenberg, J.N.S., Goldstick, J., Cevallos, W., Trostle, J. and Levy, K. (2014). Heavy rainfall events and diarrhea incidence: the role of social and environmental factors. American Journal of Epidemiology 179, 344-352.Hofstra, N. (2011). Quantifying the impact of climate change on enteric waterborne pathogen concentrations in surface water. Current Opinion in Environmental Sustainability 3, 471-479.Muirhead, R.W., Davies-Colley, R.J., Donnison, A.M. and Nagels, J.W. (2004). Faecal bacteria yields in artificial flood events: quantifying in-stream stores. Water Research 38, 1215-1224.Jamieson, R., Gordon, R., Joy, D. and Lee, H. (2004). Assessing microbial pollution of rural surface waters: a review of current watershed scale modeling approaches. Agricultural Water Management 70, 1-17.Coffey, R., Cummins, E., Bhreathnach, N., Flaherty, V.O. and Cormican, M. (2010). Development of a pathogen transport model for Irish catchments using SWAT. Agricultural Water Management 97, 101-111.Ferguson, C.M., Croke, B.F., Beatson, P.J., Ashbolt, N.J. and Deere, D.A. (2007). Development of a process-based model to predict pathogen budgets for the Sydney drinking water catchment. J Water Health 5, 187-208.Blazkova, S.D., Beven, K.J. and Smith, P.J. (2012). Transport and dispersion in large rivers: application of the aggregated dead zone model. In L. Wang and H. Garnier (eds.) System identification, environmental modelling, and control system design, 367-382. Springer London.Bencala, K.E. and Walters, R.A. (1983). Simulation of solute transport in a mountain pool-and-riffle stream: a transient storage model. Water Resources Research 19, 718-724.Romanowicz, R.J., Osuch, M. and Wallis, S. (2010). Modelling of pollutant transport in rivers under unsteady flow IAHR European Division Conference. Edinburgh, UK.Beer, T. and Young, P.C. (1983). Longitudinal dispersion in natural streams. Journal of Environmental Engineering 109, 1049-1067.Barry, D.A. and Bajracharya, K. (1996). Nonlinear reactive solute transport: A practical and fast solution method In V. Singh and B. Kumar (eds.) Water-quality hydrology (Water science and technology library, 3-17): Springer Netherlands.Yakirevich, A., Pachepsky, Y.A., Guber, A.K., Gish, T.J., Shelton, D.R. and Cho, K.H. (2013). Modeling transport of Escherichia coli in a creek during and after artificial high-flow events: Three-year study and analysis. Water Research 47, 2676-2688.Grant, S.B., Litton-Mueller, R.M. and Ahn, J.H. (2011). Measuring and modeling the flux of fecal bacteria across the sediment-water interface in a turbulent stream. Water Resources Research 47.Piorkowski, G., Jamieson, R., Bezanson, G., Truelstrup Hansen, L. and Yost, C. (2014). Reach specificity in sediment E. coli population turnover and interaction with waterborne populations. Science of The Total Environment 496, 402-413.

65

1106110711081109111011111112111311141115111611171118111911201121112211231124112511261127112811291130113111321133113411351136113711381139114011411142114311441145114611471148

Page 66: opus.bath.ac.ukopus.bath.ac.uk/53573/1/Collender_disease_risks_inla… · Web viewopus.bath.ac.uk

de Brauwere, A., de Brye, B., Servais, P., Passerat, J. and Deleersnijder, E. (2011). Modelling Escherichia coli concentrations in the tidal Scheldt river and estuary. Water Research 45, 2724-2738.de Brye, B., de Brauwere, A., Gourgue, O., Kärnä, T., Lambrechts, J., Comblen, R. and Deleersnijder, E. (2010). A finite-element, multi-scale model of the Scheldt tributaries, river, estuary and ROFI. Coastal Engineering 57, 850-863.Chapra, S.C., Pelletier, G.J. and Tao, H. (2012). QUAL2K: A modeling framework for simulating river and stream water quality. Massachussets, USA: Tufts University.Neitsch, S.L., Arnold, J.G., Kiniry, J.R. and Williams, J.R. (2011). Soil and water assessment tool theoretical documentation version 2009. Texas, USA: Texas Water Resources Institute.Tian, Y.Q., Gong, P., Radke, J.D. and Scarborough, J. (2002). Spatial and temporal modeling of microbial contaminants on grazing farmlands. Journal of Environmental Quality 31, 860-869.Droppo, I.G., Jaskot, C., Nelson, T., Milne, J. and Charlton, M. (2007). Aquaculture waste sediment stability: implications for waste migration. Water, Air, and Soil Pollution 183, 59-68.Droppo, I.G., Liss, S.N., Williams, D., Nelson, T., Jaskot, C. and Trapp, B. (2009). Dynamic existence of waterborne pathogens within river sediment compartments. Implications for water quality regulatory affairs. Environmental Science & Technology 43, 1737-1743.Fischer, H.B., List, J.E., Koh, R.C.Y., Imberger, J. and Brooks, N.H. (1979). Mixing in inland and coastal waters: Academic Press.Toprak, Z.F., Sen, Z. and Savci, M.E. (2004). Comment on "Longitudinal dispersion coefficients in natural channels". Water Research 38, 3139-3143.Toprak, Z.F. and Cigizoglu, H.K. (2008). Predicting longitudinal dispersion coefficient in natural streams by artificial intelligence methods. Hydrological Processes 22, 4106-4129.Wallis, S. and Manson, R. (2005). On the theoretical prediction of longitudinal dispersion coefficients in a compound channel. In W. Czernuszenko and P. Rowiński (eds.) Water quality hazards and dispersion of pollutants, 69-84: Springer US.Hipsey, M.R., Antenucci, J.P. and Brookes, J.D. (2008). A generic, process-based model of microbial pollution in aquatic systems. Water Resources Research 44, W07408.Leff, L.G. and Meyer, J.L. (1991). Biological availability of dissolved organic carbon along the Ogeechee River. Limnology and Oceanography 36, 315-323.Royer, T. and David, M. (2005). Export of dissolved organic carbon from agricultural streams in illinois, USA. Aquatic Sciences 67, 465-471.Tesi, T., Miserocchi, S., Acri, F., Langone, L., Boldrin, A., Hatten, J.A. and Albertazzi, S. (2013). Flood-driven transport of sediment, particulate organic matter, and nutrients from the Po River watershed to the Mediterranean Sea. Journal of Hydrology 498, 144-152.Benner, R., Opsahl, S., Chin-Leo, G., Richey, J.E. and Forsberg, B.R. (1995). Bacterial carbon metabolism in the Amazon river system. Limnology and Oceanography 40, 1262-1270.Stepanauskas, R., Laudon, H. and Jørgensen, N.O.G. (2000). High DON bioavailability in boreal streams during a spring flood. Limnology and Oceanography 45, 1298-1307.Kinner, N.E., Harvey, R.W. and Kazmierkiewicz-Tabaka, M. (1997). Effect of flagellates on free-living bacterial abundance in an organically contaminated aquifer. FEMS Microbiol Rev 20, 249-259.

66

114911501151115211531154115511561157115811591160116111621163116411651166116711681169117011711172117311741175117611771178117911801181118211831184118511861187118811891190

Page 67: opus.bath.ac.ukopus.bath.ac.uk/53573/1/Collender_disease_risks_inla… · Web viewopus.bath.ac.uk

Kinner, N.E., Harvey, R.W., Blakeslee, K., Novarino, G. and Meeker, L.D. (1998). Size-selective predation on groundwater bacteria by nanoflagellates in an organic-contaminated aquifer. Applied and Environmental Microbiology 64, 618-625.Kinner, N.E., Harvey, R.W., Shay, D.M., Metge, D.W. and Warren, A. (2002). Field evidence for a protistan role in an organically-contaminated aquifer. Environ Sci Technol 36, 4312-4318.Blanford, W.J., Brusseau, M.L., Jim Yeh, T.C., Gerba, C.P. and Harvey, R. (2005). Influence of water chemistry and travel distance on bacteriophage PRD-1 transport in a sandy aquifer. Water Res 39, 2345-2357.Pieper, A.P., Ryan, J.N., Harvey, R.W., Amy, G.L., Illangasekare, T.H. and Metge, D.W. (1997). Transport and recovery of bacteriophage PRD1 in a sand and gravel aquifer:  effect of sewage-derived organic matter. Environmental Science & Technology 31, 1163-1170.Ryan, J.N., Elimelech, M., Ard, R.A., Harvey, R.W. and Johnson, P.R. (1999). Bacteriophage PRD1 and silica colloid transport and recovery in an iron oxide-coated sand aquifer. Environmental Science and Technology 33, 63-73.Deng, M.Y. and Cliver, D.O. (1995). Antiviral effects of bacteria isolated from manure. Microb Ecol 30, 43-54.Oki, K., Yasuoka, Y. and Tamura, M. (2001). Estimation of chlorophyll-a and suspended solids concentration in rich concentration water area with remote sensing technique. Journal of the Remote Sensing Society of Japan 21, 449-457.Gameson, A.L.H. and Saxon, J.R. (1967). Field studies on effect of daylight on mortality of coliform bacteria. Water Research 1, 279-295.Ferguson, C.M., Croke, B., Ashbolt, N.J. and Deere, D.A. (2005). A deterministic model to quantify pathogen loads in drinking water catchments: pathogen budget for the Wingecarribee. Water Science and Technology 52, 191-197.Jung, A.-V., Le Cann, P., Roig, B., Thomas, O., Baurès, E. and Thomas, M.-F. (2014). Microbial contamination detection in water resources: Interest of current optical methods, trends and needs in the context of climate change. International Journal of Environmental Research and Public Health 11, 4292-4310.Tsihrintzis, V. and Hamid, R. (1997). Modeling and management of urban stormwater runoff quality: a review. Water Resources Management 11, 136-164.Page, R., Scheidler, S., Polat, E., Svoboda, P. and Huggenberger, P. (2012). Faecal indicator bacteria: groundwater dynamics and transport following precipitation and river water infiltration. Water, Air, & Soil Pollution 223, 2771-2782.Papanicolaou, A.E., M.; Krallis, G.; Prakash, S.; Edinger, J. (2008). Sediment transport modeling review—current and future developments. Journal of Hydraulic Engineering 134, 1-14.Ferguson, C., Husman, A.M.d.R., Altavilla, N., Deere, D. and Ashbolt, N. (2003). Fate and transport of surface water pathogens in watersheds. Critical Reviews in Environmental Science and Technology 33, 299-361.Elbasit, M., Salmi, A., Yasuda, H. and Ahmad, Z. (2011). Impact of rainfall microstructure on erosivity and splash soil erosion under simulated rainfall. In D. Godone (ed.) Soil erosion studies: INTECH Open Access Publisher.Merritt, W.S., Letcher, R.A. and Jakeman, A.J. (2003). A review of erosion and sediment transport models. Environmental Modelling & Software 18, 761-799.

67

1191119211931194119511961197119811991200120112021203120412051206120712081209121012111212121312141215121612171218121912201221122212231224122512261227122812291230123112321233

Page 68: opus.bath.ac.ukopus.bath.ac.uk/53573/1/Collender_disease_risks_inla… · Web viewopus.bath.ac.uk

Aksoy, H. and Kavvas, M.L. (2005). A review of hillslope and watershed scale erosion and sediment transport models. CATENA 64, 247-271.Renard, K.G., Foster, G.R., Weesies, G.A. and Porter, J.P. (1991). RUSLE: revised universal soil loss equation. Journal of Soil and Water Conservation 46, 30-33.Lane, L.J., Renard, K.G., Foster, G.R. and Laflen, J.M. (1992). Development and application of modern soil erosion prediction technology - the USDA experience. Australian Journal of Soil Research 30, 893-912.Williams, J.R., Arnold, J.G., Kiniry, J.R., Gassman, P.W. and Green, C.H. (2008). History of model development at Temple, Texas. Hydrological Sciences Journal 53, 948-960.Walker, S.E., Mostaghimi, S., Dillaha, T.A. and Woeste, F.E. (1990). Modeling animal waste management practices: impacts on bacteria levels in runoff from agricultural lands. Transactions of the ASAE 33, 807-817.Dorner, S.M., Anderson, W.B., Slawson, R.M., Kouwen, N. and Huck, P.M. (2006). Hydrologic modeling of pathogen fate and transport. Environmental Science & Technology 40, 4746-4753.Hartley, D.M. (1987). Simplified process model for water sediment yield from single storms part I - model formulation. Transactions of the ASAE 30, 710-717.Negev, M. (1967). A sediment model on a digital computer. Stanford, California, USA.Meyer, L.D. and Wischmeier, W.H. (1969). Mathematical simulation of the process of soil erosion by water. Transactions of the ASAE 12, 754-758, 762.Onstad, C.A. and Foster, G.R. (1975). Erosion modeling on a watershed. Transactions of the ASAE 18, 288-292.Bicknell, B.R., Imhoff, J.C., Kittle, J.L., Donigan, A.S. and Johanson, R.C. (1996). Hydrological simulation program - Fortran user's manual for release 11. Athens, Georgia, USA: US-EPA.Borah, D.K. and Bera, M. (2003). Watershed-scale hydrologic and nonpoint-source pollution models: review of mathematical bases. Transactions of the ASAE 46, 1553-1566.Gassman, P.W., Sadeghi, A.M. and Srinivasan, R. (2014). Applications of the SWAT model special section: overview and insights. J Environ Qual 43.Sadeghi, S.H.R., Gholami, L., Khaledi Darvishan, A. and Saeidi, P. (2014). A review of the application of the MUSLE model worldwide. Hydrological Sciences Journal 59, 365-375.Freni, G., Mannina, G. and Viviani, G. (2009). Identifiability analysis for receiving water body quality modelling. Environmental Modelling & Software 24, 54-62.Darakas, E. (2002). E. Coli kinetics - effect of temperature on the maintenance and respectively the decay phase. Environmental Monitoring & Assessment 78, 101-110.Nagels, J.W., Davies-Colley, R.J., Donnison, A.M. and Muirhead, R.W. (2002). Faecal contamination over flood events in a pastoral agricultural stream in New Zealand. Water Science and Technology 45, 45-52.Ashbolt, N.J., Grabow, W.O.K. and Snozzi, M. (2001). Indicators of microbial water quality. In L. Fewtrell and J. Bartram (eds.) Water quality: Guidelines, standards, and health: Assessment of risk and risk management for water-related infectious disease, 289-316: WHO.Jenkins, M.B., Fisher, D.S., Endale, D.M. and Adams, P. (2011). Comparative die-off of Escherichia coli 0157:H7 and fecal indicator bacteria in pond water. Environmental Science & Technology 45, 1853-1858.Schillinger, J.E. and Gannon, J.J. (1985). Bacterial adsorption and suspended particles in urban stormwater. Journal of the Water Pollution Control Federation 57, 384-389.

68

12341235123612371238123912401241124212431244124512461247124812491250125112521253125412551256125712581259126012611262126312641265126612671268126912701271127212731274127512761277

Page 69: opus.bath.ac.ukopus.bath.ac.uk/53573/1/Collender_disease_risks_inla… · Web viewopus.bath.ac.uk

Beven, K. and Binley, A. (1992). The future of distributed models: model calibration and uncertainty prediction. Hydrological Processes 6, 279-298.Bates, P.D. (2004). Remote sensing and flood inundation modelling. Hydrological Processes 18, 2593-2597.Horritt, M.S. and Bates, P.D. (2002). Evaluation of 1D and 2D numerical models for predicting river flood inundation. Journal of Hydrology 268, 87-99.Bates, P.D., Marks, K.J. and Horritt, M.S. (2003). Optimal use of high-resolution topographic data in flood inundation models. Hydrological Processes 17, 537-557.Straatsma, M.W. and Baptist, M.J. (2007). Floodplain roughness parameterization using airborne laser scanning and spectral remote sensing. Remote Sensing of Environment, 1062-1080.Smith, L.C. (1997). Sattelite remote sensing of river inundation area, stage and discharge: a review. Hydrological Processes 11, 1427-1439.Patro, S., Chatterjee, C., Mohanty, S., Singh, R. and Raghuwanshi, N.S. (2009). Flood inundation modeling using MIKE flood and remote sensing data. Journal of the Indian Society of Remote Sensing 37, 107-118.Bates, P.D., Horritt, M.S., Smith, C.N. and Mason, D. (1997). Integrating remote sensing observations of flood hydrology and hydraulic modelling. Hydrological Processes 11, 1777-1795.Bates, P.D. (2012). Integrating remote sensing data with flood inundation models: how far have we got? Hydrological Processes 26, 2515-2521.Di Baldassarre, G. and Uhlenbrook, S. (2012). Is the current flood of data enough? A treatise on research needs for the improvement of flood modelling. Hydrological Processes 26, 153-158.Bradford, S.A. and Schijven, J. (2002). Release of Cryptosporidium and Giardia from dairy calf manure: impact of solution salinity. Environmental Science & Technology 36, 3916-3923.Schijven, J.F., Bradford, S.A. and Yang, S. (2004). Release of Cryptosporidium and Giardia from dairy cattle manure: physical factors. J Environ Qual 33, 1499-1508.Wood, M., Hostache, R., Neal, J., Wagener, T., Giustarini, L., Chini, M., Corato, G., Matgen, P. and Bates, P. (2016). Calibration of channel depth and friction parameters in the LISFLOOD-FP hydraulic model using medium resolution SAR data. Hydrol. Earth Syst. Sci. Discuss. 2016, 1-24.Beven, K. (2010b). Distributed models and uncertainty in flood risk management. In Flood risk science and management, 289-312: Wiley-Blackwell.Svensson, C., Kjeldsen, T.R. and Jones, D.A. (2013). Flood frequency estimation using a joint probability approach within a Monte Carlo framework. Hydrological Sciences Journal 58, 8-27.Rode, M., Arhonditsis, G., Balin, D., Kebede, T., Krysanova, V., van Griensven, A. and van der Zee, S. (2010). New challenges in integrated water quality modelling. Hydrological Processes 24, 3447-3461.

69

12781279128012811282128312841285128612871288128912901291129212931294129512961297129812991300130113021303130413051306130713081309131013111312131313141315