Hyporheic Zone in Urban Streams: A Review and Opportunities for Enhancing Water Quality and...

22
Hyporheic Zone in Urban Streams: A Review and Opportunities for Enhancing Water Quality and Improving Aquatic Habitat by Active Management Justin E. Lawrence, 1,2, * Magnus E. Skold, 1,3 Fatima A. Hussain, 1,4 David R. Silverman, 1,3 Vincent H. Resh, 1,2 David L. Sedlak, 1,5 Richard G. Luthy, 1,4 and John E. McCray 1,3 1 Engineering Research Center for Re-Inventing the Nation’s Urban Water Infrastructure (ReNUWIt), National Science Foundation, Stanford, California. Departments of 2 Environmental Science, Policy, & Management and 5 Civil & Environmental Engineering, University of California, Berkeley, California. 3 Department of Civil & Environmental Engineering, Colorado School of Mines, Golden, Colorado. 4 Department of Civil & Environmental Engineering, Stanford University, Stanford, California. Received: June 15, 2012 Accepted in revised form: May 6, 2013 Abstract Tremendous opportunities exist for enhancing water quality and improving aquatic habitat by actively managing urban water infrastructure to operate in conjunction with natural systems. The hyporheic zone (HZ) of streams, which is the area of active mixing between surface water and groundwater, is one such system that is overlooked by many water professionals, because the state of the science on this topic has not been transferred into practice. As a biogeochemically active zone, the HZ offers great potential to provide natural treatment of organic compounds, nutrients, and pathogens in urban streams, which are often strongly impacted by flow modifications and water pollution. Reliable treatment is most likely in streams in which the majority of flow occurs through the HZ, the flow is aerated, and sufficient residence times occur, which may be limited to specific channel morphologies and seasons. Integration of the HZ into stream management plans could also provide quality habitat in a landscape with increasingly depauperate biodiversity. Here, we review current knowledge on hydrological, chemical, and biological aspects of the HZ, with a focus on urban settings, and include a set of examples drawn from the literature of low-flow, effluent-dominated streams in which there is significant hyporheic flow and potential for contaminant attenuation. The HZ can be incorporated much more effectively into urban water manage- ment, including stream restoration efforts, by understanding the surface and subsurface features conducive to HZ flow and the water-quality and biodiversity improvements that can be gained in the HZ without posing unreasonable risk. The main barriers to implementation of HZ considerations include lack of information, absence of established metrics for evaluating success, small number of controlled HZ experiments in urban settings, and concern over risks to both public health and aquatic organisms. A combination of field studies, laboratory experiments, and model development that consider hydrological, chemical, and biological inter- actions in the HZ can overcome these barriers. Key words: biodiversity; dry season flows; micropollutants; natural systems enhancement; nutrients; subsurface flow; urban water infrastructure; wastewater effluent Introduction U rban streams and rivers have a significant role to play in the reinvention of urban water infrastructure. For example, through implementation of appropriate manage- ment tools, it may be possible to create systems in which streams flowing out of cities are cleaner than streams flowing into them (Daigger, 2007), and key components of the natural flow regime are nearly intact (Poff et al., 2010). Urban stream channels are commonly incised, lowering the ground water level, which can reduce base flow and accelerate channel erosion and sediment production. They also sometimes have impermeable channel linings, which reduces interaction be- tween surface and ground waters. Hydromodification of stream channels, runoff from impervious surfaces, and input of low quality, wastewater effluents are some of the many variables associated with urban growth and development *Corresponding author: Department of Environmental Science, Pol- icy, and Management, University of California, Berkeley, CA 94720- 3114. Phone: 510-642-6315; Fax: 510-642-7428; E-mail: jlawrence@ berkeley.edu ENVIRONMENTAL ENGINEERING SCIENCE Volume 30, Number 8, 2013 ª Mary Ann Liebert, Inc. DOI: 10.1089/ees.2012.0235 480

Transcript of Hyporheic Zone in Urban Streams: A Review and Opportunities for Enhancing Water Quality and...

Hyporheic Zone in Urban Streams:A Review and Opportunities for Enhancing Water Quality

and Improving Aquatic Habitat by Active Management

Justin E. Lawrence,1,2,* Magnus E. Skold,1,3 Fatima A. Hussain,1,4 David R. Silverman,1,3

Vincent H. Resh,1,2 David L. Sedlak,1,5 Richard G. Luthy,1,4 and John E. McCray1,3

1Engineering Research Center for Re-Inventing the Nation’s Urban Water Infrastructure (ReNUWIt),

National Science Foundation, Stanford, California.

Departments of 2Environmental Science, Policy, & Management and 5Civil & Environmental

Engineering, University of California, Berkeley, California.3Department of Civil & Environmental Engineering, Colorado School of Mines, Golden, Colorado.

4Department of Civil & Environmental Engineering, Stanford University, Stanford, California.

Received: June 15, 2012 Accepted in revised form: May 6, 2013

Abstract

Tremendous opportunities exist for enhancing water quality and improving aquatic habitat by actively managingurban water infrastructure to operate in conjunction with natural systems. The hyporheic zone (HZ) of streams,which is the area of active mixing between surface water and groundwater, is one such system that is overlookedby many water professionals, because the state of the science on this topic has not been transferred into practice. Asa biogeochemically active zone, the HZ offers great potential to provide natural treatment of organic compounds,nutrients, and pathogens in urban streams, which are often strongly impacted by flow modifications and waterpollution. Reliable treatment is most likely in streams in which the majority of flow occurs through the HZ, the flowis aerated, and sufficient residence times occur, which may be limited to specific channel morphologies and seasons.Integration of the HZ into stream management plans could also provide quality habitat in a landscape withincreasingly depauperate biodiversity. Here, we review current knowledge on hydrological, chemical, andbiological aspects of the HZ, with a focus on urban settings, and include a set of examples drawn from theliterature of low-flow, effluent-dominated streams in which there is significant hyporheic flow and potentialfor contaminant attenuation. The HZ can be incorporated much more effectively into urban water manage-ment, including stream restoration efforts, by understanding the surface and subsurface features conducive toHZ flow and the water-quality and biodiversity improvements that can be gained in the HZ without posingunreasonable risk. The main barriers to implementation of HZ considerations include lack of information,absence of established metrics for evaluating success, small number of controlled HZ experiments in urbansettings, and concern over risks to both public health and aquatic organisms. A combination of field studies,laboratory experiments, and model development that consider hydrological, chemical, and biological inter-actions in the HZ can overcome these barriers.

Key words: biodiversity; dry season flows; micropollutants; natural systems enhancement; nutrients; subsurfaceflow; urban water infrastructure; wastewater effluent

Introduction

Urban streams and rivers have a significant role to playin the reinvention of urban water infrastructure. For

example, through implementation of appropriate manage-ment tools, it may be possible to create systems in which

streams flowing out of cities are cleaner than streams flowinginto them (Daigger, 2007), and key components of the naturalflow regime are nearly intact (Poff et al., 2010). Urban streamchannels are commonly incised, lowering the ground waterlevel, which can reduce base flow and accelerate channelerosion and sediment production. They also sometimes haveimpermeable channel linings, which reduces interaction be-tween surface and ground waters. Hydromodification ofstream channels, runoff from impervious surfaces, and inputof low quality, wastewater effluents are some of the manyvariables associated with urban growth and development

*Corresponding author: Department of Environmental Science, Pol-icy, and Management, University of California, Berkeley, CA 94720-3114. Phone: 510-642-6315; Fax: 510-642-7428; E-mail: [email protected]

ENVIRONMENTAL ENGINEERING SCIENCEVolume 30, Number 8, 2013ª Mary Ann Liebert, Inc.DOI: 10.1089/ees.2012.0235

480

that impair urban streams, and these stresses are expected tointensify in the future with additional population growth andclimate change (Grischek et al., 2002).

To restore the ecology and aesthetics of urban streams,municipal wastewater effluent and urban runoff should beenvisioned as resources, rather than wastes (Bischel et al., 2013).However, contaminants in these water sources have potentialto affect drinking-water supplies and degrade aquatic life. Forexample, pharmaceuticals (e.g., acetaminophen), personal-careproducts (e.g., triclosan), and consumer products (e.g., per-fluorinated surfactants) can reach aquifers through the hy-porheic zone (HZ). In the HZ, surface water and groundwateris exchanged, chemical transformations of various types occur,and microorganisms and animals live (Boulton et al., 2010;Bianchin et al., 2011). The HZ is increasingly recognized for itspotential to attenuate contaminants (Biksey and Gross, 2001).As a result of this function, it has been metaphorically called theriver’s liver (Fischer et al., 2005).

The HZ provides interrelated ecosystem services related towater quality and aquatic habitat. For example, transforma-tions of redox-active chemicals occur in the HZ, which arepartly facilitated by the suitable conditions that an intact HZprovides (Boano et al., 2010a). When relatively long residencetimes occur in HZs (e.g., hours to days), significant attenuationof contaminants and buffering of water temperature can oc-cur, which can be considered a valuable ecosystem-servicebenefit associated with their management and preservation.Furthermore, microorganisms and animals use the HZ as botha primary and a refuge habitat at different life stages, createhydrologic flow paths that enhance permeability, and alsohave the potential to remove nutrients and trace organicchemicals (Clinton et al., 2010; Wood et al., 2010).

Most studies conducted on the HZ have taken place eitherin pristine headwaters or in agricultural areas rather than inurban areas, which exhibit site-specific dynamics that may notnecessarily be generalized to all streams (Malcolm et al., 2010;Lewandowski et al., 2011). Urban streams are often located inlowlands (where cities are) and can have less turbulent,though higher-velocity (as a result of channelization for floodcontrol) and higher-volume (as a result of flashiness fromimpervious surfaces) flows than mountain streams, and theyoften receive elevated sediment inputs and have reducedcanopy coverage as a result of human development (Walshet al., 2005). Furthermore, the turbulence level would alsodepend on whether the stream is located in a forest (and thusmore likely to be composed of riffle-steps-pools) or in mead-ows (smooth meanders). The discharge of urban streams thatare small or which exist in arid or semi-arid climates may bedominated by effluent from municipal wastewater treatmentplants (WWTPs). In streams receiving municipal wastewatereffluent, the HZ may contain relatively high concentrations ofnutrients and trace organic contaminants, as well as highwater temperatures relative to what is natural for the region(Tufenkji et al., 2002).

The objective of this article is to review the processes in theHZ that influence the fate of nutrients and trace organic con-taminants, as well as the structure and function of ecologicalcommunities in the HZ, and to suggest practical HZ manage-ment strategies for enhancing both water quality and aquaticbiodiversity. We identify (1) gaps in the current understandingof hyporheic processes in urban streams; (2) research needs forachieving contaminant attenuation and ecological improve-

ment; and (3) barriers to the implementation of HZ manage-ment projects. The concepts examined are illustrated with acollection of examples, primarily drawn from low-flow,effluent-dominated streams in urban areas in the westernUnited States and other developed countries in which thereis high population growth and increasing water-demandpressures.

Discussion

Hydrology and the HZ

Many variables have been used to define the HZ, includingthose based on a fixed distance within or from the stream (e.g.,a prescribed depth), ecology (e.g., shifts in floral and faunaldistributions), morphology (e.g., changes in substrate type),chemistry, hydrology, residence time, and combinations ofthese variables (Williams, 1989; Boulton et al., 2010; Gooseff,2010; Bianchin et al., 2011). The definition selected can dra-matically affect the defined size of the HZ (White, 1993). Forexample, the depth of the HZ can range from several centi-meters when bedrock is close to the surface to > 10 m whenstreams flow over a sand or alluvial substrate (Chen, 2011;Stelzer et al., 2011). Moreover, the lateral extent of the HZcan range from just a few centimeters in bedrock-confinedchannels to multiple kilometers in large river systems withwide alluvial floodplains (Lapworth et al., 2009; Sawyer et al.,2009).

In this article, we define the HZ using the hydrologicaldefinition in which flowpaths originate and terminate at thestream, including the concept of residence time (Gooseff,2010). For example, a 24-h HZ would refer to a region inwhich hyporheic water takes 24 h to pass from the streamthrough the subsurface zone of active mixing of surfacewater and ground water and then back to the stream. We usethe term hyporheic exchange in the sense of Harvey et al.(1996) to refer to exchanges of water between channels andthe subsurface at small scales (centimeter to meter). The sizeof the HZ under this hydrological definition can be measuredin the field using tracers (such as NaCl, bromide, or tem-perature), and it fluctuates through space and time as thereservoirs of surface water and groundwater expand andcontract. High-resolution distributed temperature sensingcan be used to quantify spatiotemporal variability in verticalhyporheic flux (Briggs et al., 2012). The residence-time con-cept allows efficient linkage of the HZ to chemical transfor-mation (Fig. 1).

A selection of case studies of low-flow, effluent-dominatedurban streams in which there is documented pollutantattenuation and the HZ is likely a key factor was made toillustrate the range of settings, hydrology, pollution, attenu-ation, and implications for HZ management that are likely tobe encountered in urban areas (Table 1), and additional ex-amples are provided throughout the text to illustrate thatother sites can exhibit similar characteristics. In Upper SilverCreek and Coyote Creek (Table 1), for example, residencetimes in the HZ were observed to range from 15 to 40 minalong the *5 km reach examined. Under these conditions ofshort residence time, little degradation of dissolved trace or-ganic contaminants was expected. In contrast, confirmation ofsignificant natural attenuation of contaminants combinedwith low flows, and presumably long residence times, wasobserved in the Santa Cruz River (Table 1).

ACTIVE MANAGEMENT OF HYPORHEIC ZONES IN URBAN STREAMS 481

Hydrology has been widely recognized to be the domi-nant driver of ecological structure and function in streamecosystems (Resh et al., 1988; Power et al., 1995), and itcertainly affects processes in the HZ (Boulton et al., 2010;Robertson and Wood, 2010). The connection of the HZ withthe stream is best thought of conceptually in four dimen-sions, that is, the three spatial dimensions and time (Ward,1989). As with stream management that is based on thenatural flow regime principle (Richter et al., 1997; Poff et al.,1997; Tharme, 2003), the magnitude, frequency, duration,timing, and rate of change of flows through the HZ are allimportant components that should be considered alongwith water quality for a management program to be trulyeffective in meeting a stream’s ecological needs (Hawkinset al., 2010; Poff et al., 2010).

The hyporheic flow regime can be altered by a variety ofhuman activities, including those that change the streamflowregime or alter the streambed surface or subsurface (Boanoet al., 2010b; Maier and Howard, 2011). For example, thestreamflow regime in the Salt and Gila Rivers is altered byelevated storm runoff from impervious surfaces, upstreamdam releases, and wastewater effluent discharge (Table 1).

The associated reduction in retention time in many urbanstreams (because urban runoff is associated with high veloc-ities and high volumes) can result in reduced infiltration intothe HZ, giving rise to lower relative volumes and residencetimes (Grimm et al., 2005). Channelization, culverting, ar-moring of banks with riprap, and sediment inputs are somecommon activities in urban streams that would also alter thenatural flow regime from surface water to the HZ (Brownet al., 2009; Stranko et al., 2012). Finally, the hydrology of astream has a major impact on the processes in the associatedHZ that affect water quality, including the establishment ofappropriate redox conditions (Lewandowski et al., 2011).

Water quality and the HZ

The HZ forms a unique and dynamic ecosystem that is bothinfluenced by and can significantly influence neighboringstream-water and groundwater quality. The cycling of ele-ments, organic compounds, nitrogen (N), and phosphorous(P) in the HZ has been studied in depth (Findlay et al., 1993;Crenshaw et al., 2010; Lapworth et al., 2011). An under-standing of the fate and transport of these solutes, which is

FIG. 1. Simplified conceptual overview of: (A) hyporheic zone (HZ) management in urban streams encompassing thehydrological, ecological, and water quality considerations associated with wastewater effluent in urban streams and in-cluding the concept of residence time (RT); and (B) microbially mediated pathways for transformation of nutrients andbiodegradation of trace organic contaminants (TrOCs). This figure does not cover all possible reactions in the HZ.

482 LAWRENCE ET AL.

Ta

bl

e1.

Ca

se

St

ud

ie

so

fL

ow

-F

lo

w,

Effl

ue

nt

-D

om

in

at

ed

Ur

ba

nS

tr

ea

ms

Wh

er

eT

he

re

Is

Do

cu

me

nt

ed

Po

ll

ut

an

tA

tt

en

ua

tio

n

an

dt

he

Hy

po

rh

eic

Zo

ne

Is

Lik

el

ya

Ke

yF

ac

to

r

Sit

eS

etti

ng

and

hy

dro

log

yP

ollu

tion

and

atte

nu

atio

nIm

pli

cati

ons

for

HZ

man

agem

ent

Sal

tan

dG

ila

Riv

ers

(SG

R),

Ph

oen

ix,

AZ

Flo

win

gth

rou

gh

the

city

of

Ph

oen

ix(3

3cm

/y

ear

mea

nra

infa

ll),

the

SG

Rd

rain

aw

ater

shed

of

32,0

00k

m2.

Th

eyar

en

atu

rall

yep

hem

eral

and

flas

hy

,fl

ow

ing

inre

spo

nse

top

reci

pit

atio

n,

up

stre

amd

amre

leas

es,

and

WW

TP

effl

uen

t.a,b

Th

e20

09M

AF

of

the

Sal

tR

iver

at51

stA

ven

ue

inP

ho

enix

was

1.4

m3/

sM

DF

oft

enze

ro,

wh

ile

the

2011

MA

Ffl

ow

of

the

Gil

aR

iver

asit

exit

sth

eci

tyw

as0.

094

m3/

sM

DF

oft

enze

ro(U

SG

Sg

aug

ing

stat

ion

s#

0951

2406

and

0951

4100

,re

spec

tiv

ely

).

Wit

hth

eex

cep

tio

no

fse

aso

nal

flo

od

per

iod

s,fl

ow

isp

rim

aril

yef

flu

ent

dis

char

ged

fro

m18

WW

TP

s.a

Ast

ud

yo

ffi

ve

stre

ams

inth

eP

ho

enix

Met

roA

rea

rev

eale

dd

ecre

ased

Nre

ten

tio

nan

dh

yp

orh

eic

exch

ang

ein

thes

est

ream

sco

mp

ared

wit

hu

nd

istu

rbed

stre

ams

insi

mil

arg

eog

rap

hic

sett

ing

s.b

,cR

elat

ive

HZ

size

was

defi

ned

by

the

exte

nt

of

tran

sien

tst

ora

ge

and

rate

of

gro

un

d-

wat

er-

surf

ace

wat

erex

chan

ge

asm

easu

red

by

tem

po

ral

pat

tern

sin

Br

-tr

acer

test

s.c

Th

eS

GR

are

typ

ical

of

urb

anst

ream

sin

the

arid

and

sem

iari

dw

est

that

are

rece

ivin

gW

WT

Pef

flu

ent.

dA

ssu

ch,

HZ

man

agem

ent

too

lsd

evel

-o

ped

for

thes

est

ream

sco

uld

be

wid

ely

app

lica

-b

le.

Th

eS

GR

and

thei

rn

um

ero

us

trib

uta

ries

are

wel

lch

arac

teri

zed

and

hav

ev

ary

ing

lev

els

of

urb

anm

od

ifica

tio

n,

wh

ich

pro

vid

esan

idea

lse

ttin

gto

stu

dy

the

role

of

HZ

pro

cess

esin

WW

DC

atte

nu

atio

n.

San

taC

ruz

Riv

er(S

CR

),N

og

ales

,A

ZT

he

SC

Rd

rain

sa

3,00

0-k

m2

allu

via

lb

asin

(40

cm/

yea

rm

ean

rain

fall

)at

ap

oin

tlo

cate

d16

km

bel

ow

NIW

TP

.eIt

isep

hem

eral

alo

ng

mo

sto

fit

sle

ng

th,b

ut

isp

eren

nia

lb

elo

wN

IWT

P.e

Th

eM

AF

(199

6–20

11)

was

0.84

m3/

sw

ith

MD

Fas

low

asze

ro(U

SG

Sg

aug

ing

stat

ion

#09

4817

40).

Th

eS

CR

iso

ften

100%

effl

uen

tb

elo

wth

eN

IWT

P,

wh

ich

dis

char

ges

*15

MG

D.d

,fN

atu

ral

atte

nu

-at

ion

of

WW

DC

and

nu

trie

nts

inth

eri

ver

bel

ow

NIW

TP

has

bee

nco

nfi

rmed

,d

esp

ite

the

form

a-ti

on

of

acl

og

gin

gla

yer

asso

ciat

edw

ith

WW

TP

effl

uen

t,w

hic

hd

ecre

ases

stre

amb

edh

yd

rau

lic

con

du

ctiv

ity

and

inh

ibit

sh

yp

orh

eic

exch

ang

e.e,f

Her

e,th

eH

Zw

asd

efin

edq

ual

itat

ivel

yas

azo

ne

bel

ow

the

shal

low

stre

amb

edth

ath

ost

sst

ream

-aq

uif

erin

tera

ctio

ns.

Th

eco

nfi

rmat

ion

of

nat

ura

lat

ten

uat

ion

of

WW

DC

com

bin

edw

ith

low

flo

ws

sug

ges

tsth

atth

eH

Zm

ayp

lay

ala

rge

role

inp

oll

uta

nt

tran

sfo

rmat

ion

inth

isri

ver

.C

log

gin

gla

yer

sas

soci

ated

wit

hW

WT

Pef

flu

ent

wer

ere

mo

ved

du

rin

gfl

oo

dfl

ow

s,su

gg

esti

ng

that

eng

inee

red

hig

h-fl

ow

re-

leas

esco

uld

be

am

anag

emen

tst

rate

gy

for

pro

mo

tin

gh

yp

orh

eic

exch

ang

ein

sim

ilar

stre

ams.

San

taA

na

Riv

er(S

AR

),O

ran

ge

Co

un

ty,

CA

Th

eS

AR

dra

ins

a7,

250

km

2w

ater

shed

(25

to50

cm/

yea

rra

infa

ll)

and

isan

imp

ort

ant

sou

rce

of

aqu

ifer

rech

arg

efo

rth

ere

gio

n.g

Wit

ha

nat

ura

lst

ream

bed

and

eng

inee

red

sid

ewal

lsal

on

ga

po

rtio

no

fit

sle

ng

th,

flo

win

the

SA

Rca

nco

nsi

sto

f10

0%ef

flu

ent

fro

ma

loca

lW

WT

Pd

uri

ng

the

dry

seas

on

.h,i

MA

F(1

941–

2011

)in

the

SA

Rb

elo

wP

rad

oD

amw

as7

m3/

sw

ith

MD

Fas

low

as0.

07m

3/

s(U

SG

Sg

aug

ing

stat

ion

#11

0740

00).

Sev

eral

stu

die

sh

ave

iden

tifi

edth

ep

rese

nce

and

nat

ura

lat

ten

uat

ion

of

WW

DC

inth

eS

AR

.h,i

,j

Ph

arm

aceu

tica

lsan

dth

eir

met

abo

lite

sw

ere

at-

ten

uat

edsu

bst

anti

ally

on

a12

-km

reac

ho

fth

eS

AR

,w

ith

bio

tran

sfo

rmat

ion

and

sorp

tio

np

lay

-in

gp

rim

ary

role

s.h

,jIn

dic

ato

rco

mp

ou

nd

s,in

-cl

ud

ing

,N

TA

,A

PE

Cs,

bro

mo

chio

ro-

and

dib

rom

oac

etic

acid

s,w

ere

sig

nifi

can

tly

atte

nu

-at

edas

they

infi

ltra

ted

thro

ug

hth

eH

Z.i

Her

e,th

eH

Zw

asd

efin

edas

the

zon

eo

fp

erco

lati

on

bet

wee

nth

eri

ver

bed

and

am

on

ito

rin

gw

ell

9m

dee

pan

dd

irec

tly

adja

cen

tto

SA

R.

Th

eS

AR

isw

ell

suit

edto

serv

eas

ate

stb

edfo

rH

Zm

anag

emen

tp

ract

ices

,b

ecau

seit

isb

oth

wel

lch

arac

teri

zed

hy

dro

geo

log

ical

lyan

dh

asd

em-

on

stra

ted

nat

ura

lat

ten

uat

ion

of

WW

DC

inH

Zp

erm

eate

.

Up

per

Sil

ver

Cre

eks

(US

C)

and

Co

yo

teC

reek

s(C

C),

San

Jose

,C

A

US

Can

dC

Cd

rain

a48

-km

2ca

tch

men

t(4

5cm

/y

ear

mea

nra

infa

ll).

kU

SC

has

MD

Fas

low

as0.

04m

3/

s,th

ou

gh

this

can

incr

ease

by

afa

cto

r10

0–10

00d

uri

ng

rain

even

ts.k

MA

Fd

ata

are

un

avai

lab

leat

US

C.

MA

F(1

999–

2010

)in

CC

atM

ilp

itas

,C

A,

was

1.3

m3/

sw

ith

MD

Fas

low

as0.

2m

3/

s(U

SG

Sg

aug

ing

stat

ion

#11

1721

7).

US

Can

dC

Cd

on

ot

curr

entl

yre

ceiv

eW

WT

Pef

flu

ent,

bu

tm

icro

po

llu

tan

tsar

est

ill

det

ecte

din

the

riv

ersy

stem

.k,l

PF

Cs

wer

ed

etec

ted

insu

rfac

ew

ater

,h

yp

orh

eic

wat

er,

and

the

un

der

lyin

gaq

uif

er,

and

neg

lig

ible

atte

nu

atio

nw

aso

b-

serv

ed.k

,lIt

was

hy

po

thes

ized

that

the

stu

dy

reac

hw

asto

osh

ort

tod

etec

tat

ten

uat

ion

.k,l

Th

eH

Zw

asta

ken

asth

ezo

ne

imm

edia

tely

un

der

-ly

ing

the

riv

erb

edw

her

ew

ater

flo

wed

bet

wee

nth

est

ream

and

sub

surf

ace

wit

hre

sid

ence

tim

esin

the

ord

ero

f15

–60

min

asm

easu

red

by

tim

ese

ries

of

riv

eran

dg

rou

nd

wat

erte

mp

erat

ure

.k,l

Rec

alci

tran

tP

FC

sar

ere

adil

ytr

ansp

ort

edfr

om

surf

ace

wat

erin

tou

nd

erly

ing

aqu

ifer

s,k

,lb

ut

the

exte

nt

tow

hic

hth

eyar

eat

ten

uat

edin

the

HZ

rem

ain

sto

be

det

erm

ined

.U

SC

and

CC

may

serv

eas

ah

igh

lysu

itab

lete

stb

edfo

rH

Zat

ten

-u

atio

nst

ud

ies

bec

ause

of

the

var

ied

stre

amb

edm

orp

ho

log

yan

dth

ed

ocu

men

ted

mic

rop

oll

uta

nt

con

tam

inat

ion

.T

he

San

taC

lara

Val

ley

Wat

erD

istr

ict

(San

Jose

,C

A)

con

sid

ered

aug

men

tin

gU

SC

and

CC

flo

ww

ith

tert

iary

trea

ted

was

te-

wat

er,

bu

tth

ep

roje

ctw

asd

isco

nti

nu

edd

ue

toco

nce

rns

ov

erW

WD

C.k

,l

(con

tin

ued

)

483

Ta

bl

e1.

(Co

nt

in

ue

d)

Sit

eS

etti

ng

and

hy

dro

log

yP

ollu

tion

and

atte

nu

atio

nIm

pli

cati

ons

for

HZ

man

agem

ent

Bo

uld

erC

reek

(BC

),B

ou

lder

,C

OB

Cd

rain

sa

1,16

0-k

m2

wat

ersh

ed(5

5cm

/y

ear

mea

nra

infa

ll)

and

isp

eren

nia

l.m

Inth

eci

tyo

fB

ou

lder

,W

WT

Pef

flu

ent

can

acco

un

tfo

rap

-p

rox

imat

ely

77%

of

the

stre

amfl

ow

.M

AF

inB

Cb

elo

wth

e75

thS

tree

tW

WT

P(1

984–

2011

)w

as2.

6m

3/

sw

ith

MD

Fas

low

as0.

025

m3/

sd

uri

ng

the

dry

seas

on

(US

GS

gau

gin

gst

atio

n#

0673

0200

).T

he

75th

Str

eet

WW

TP

dis

char

ges

on

aver

age*

0.65

m3/

so

fef

flu

ent.

m

41W

WD

Ch

ave

bee

nd

etec

ted

inB

C.m

,n,o

Th

een

do

crin

e-d

isru

pti

ng

det

erg

ent

com

po

un

d4-

NP

was

com

ple

tely

min

eral

ized

ino

xic

mic

roco

sms

pre

par

edfr

om

BC

sed

imen

tfr

om

abo

ve

the

WW

TP

,w

hil

em

iner

aliz

atio

nw

assi

gn

ifica

ntl

yd

imin

ish

edin

ox

icm

icro

cosm

sp

rep

ared

wit

hB

Cse

dim

ent

fro

mb

elo

wth

eW

WT

P.p

Nei

ther

mic

roco

sms

pre

par

edw

ith

sed

imen

tsfr

om

abo

ve

no

rb

elo

wth

eW

WT

Psh

ow

edsi

gn

ifica

nt

min

eral

izat

ion

of

4-N

Pw

hen

incu

bat

edu

nd

eran

ox

icco

nd

itio

ns.

pT

he

HZ

was

no

td

irec

tly

stu

die

d.

BC

can

serv

eas

anil

lust

rati

ve

con

tras

tto

eph

em-

eral

stre

ams

inth

ear

idW

est

that

rece

ive

WW

TP

effl

uen

t.H

Zm

anag

emen

tp

ract

ices

for

ast

ream

sim

ilar

toB

Cw

ith

turb

ule

nt

flo

wan

dm

any

recr

eati

on

alu

sers

du

rin

gth

esu

mm

erm

ayd

iffe

rg

reat

lyfr

om

tho

seo

fan

eph

emer

alar

id-l

and

stre

am.

BC

sed

imen

th

assh

ow

nth

eca

pac

ity

toae

rob

ical

lyb

iod

egra

de

mic

rop

oll

uta

nts

,su

chas

4-N

P;

ho

wev

er,

incr

ease

dB

OD

and

low

erD

Ofr

om

WW

TP

effl

uen

td

ecre

ases

effi

cien

cy.p

So

uth

Pla

tte

Riv

er(S

PR

),D

env

er,

CO

Th

eS

PR

dra

ins

ab

asin

of

63,0

00k

m2

(40

cm/

yea

rm

ean

rain

fall

)w

ith

>10

0W

WT

Ps.

qT

he

stre

am-

bed

con

sist

so

fw

ell-

sort

ed,

med

ium

-gra

ined

san

d,

and

hy

po

rhei

cex

chan

ge

isfa

st.r,

sM

AF

(198

3–20

11)

inth

eS

PR

bel

ow

the

city

of

Den

ver

was

7.4

m3/

sw

ith

MD

Fas

low

as0.

34m

3/

s(U

SG

Sg

aug

ing

#06

7115

65).

WW

TP

effl

uen

tac

cou

nts

for

>95

%o

fb

asefl

ow

.q,t

Ho

url

yv

aria

-ti

on

sin

WW

TP

dis

char

ge

rate

sca

use

mea

sura

ble

var

iati

on

sin

riv

erst

age

for

>35

km

do

wn

-st

ream

.t

Ho

url

yfl

uct

uat

ion

sin

effl

uen

td

isch

arg

era

tes

fro

mD

env

er’s

larg

est

WW

TP

dir

ectl

yaf

fect

stre

amst

age,

and

hen

ceh

yp

orh

eic

exch

ang

eo

na

60-k

mst

retc

hd

ow

nst

ream

.tH

igh

erh

yp

orh

eic

exch

ang

era

tes

led

toin

crea

sed

up

tak

eo

fD

Oin

the

HZ

.t

‘‘Hy

po

rhei

cex

chan

ge’

’w

asd

efin

edas

flo

wac

ross

the

sed

imen

t-w

ater

inte

rfac

ean

dw

asd

eter

min

edfr

om

the

ver

tica

lh

yd

rau

lic

gra

die

nt

(bet

wee

nsu

rfac

ew

ater

and

dep

tho

f30

cm)

and

hy

dra

uli

cco

nd

uct

ivit

yo

fb

ott

om

sed

imen

ts.t

DO

sam

ple

sin

the

HZ

wer

eta

ken

ata

dep

tho

f30

cm.t

DO

up

tak

era

tes

by

stre

amb

edse

dim

ent

wer

e2–

5ti

mes

hig

her

than

that

req

uir

edb

ym

od

els

wh

ich

sim

ula

teS

PR

DO

dy

nam

ics,

sug

ges

tin

ga

mo

resi

gn

ifica

nt

role

of

the

HZ

inco

ntr

oll

ing

riv

erD

Od

yn

amic

sth

anp

rev

iou

sly

tho

ug

ht.

tO

ther

stu

die

sh

ave

do

cum

ente

dth

eb

iod

egra

dat

ion

of

end

ocr

ine-

dis

rup

tin

gch

emi-

cals

inS

PR

sed

imen

tm

icro

cosm

s.p

,uT

hes

est

ud

ies

sho

wed

that

4-N

P,

17b

-est

rad

iol,

estr

on

e,an

dte

sto

ster

on

ew

ere

atte

nu

ated

tov

ary

ing

deg

rees

inm

icro

cosm

sp

rep

ared

fro

mS

PR

sed

-im

ent

fro

mab

ov

ean

db

elo

wth

eW

WT

Pw

ith

bio

deg

rad

atio

np

lay

ing

asi

gn

ifica

nt

role

.

Wat

eru

tili

ties

wer

eu

nk

no

win

gly

affe

ctin

gh

yp

or-

hei

cex

chan

ge

pro

cess

esw

ith

ho

url

yfl

uct

uat

ion

so

fW

WT

Pef

flu

ent

dis

char

ge

rate

s.tU

nin

ten

tio

nal

anth

rop

og

enic

effe

cts

on

HZ

pro

cess

essh

ou

ldb

eb

ette

rch

arac

teri

zed

toin

tell

igen

tly

man

age

thes

eco

mp

lex

syst

ems.

As

we

con

tem

pla

teH

Zm

an-

agem

ent,

we

sho

uld

reco

gn

ize

that

man

agem

ent

pra

ctic

esm

ayh

ave

con

flic

tin

gre

sult

s.F

or

in-

stan

ce,

effo

rts

toim

pro

ve

DO

dy

nam

ics

inef

flu

ent-

do

min

ated

riv

ers

mig

ht

con

sid

erre

du

c-in

gh

yp

orh

eic

exch

ang

e.t

Wh

ile

red

uci

ng

hy

po

r-h

eic

exch

ang

em

ayp

rov

ide

ad

esir

edo

utc

om

efo

rD

Od

yn

amic

s,re

du

ced

exch

ang

ew

ou

ldre

sult

inu

nfa

vo

rab

leco

nd

itio

ns

for

the

atte

nu

a-ti

on

of

man

ym

icro

po

llu

tan

ts.

Up

per

Wil

lam

ette

Riv

er(U

WR

),C

orv

alli

s,O

R

Th

eU

WR

dra

ins

ab

asin

of

31,0

00-k

m2

(114

cm/

yea

rm

ean

rain

fall

)an

dfl

ow

so

ver

ap

oro

us

sub

stra

teth

atis

con

du

civ

eto

hy

po

rhei

cex

-ch

ang

e.v

Th

eC

orv

alli

sW

WT

Pd

isch

arg

es8–

10M

GD

into

the

riv

ery

ear

rou

nd

.wT

he

MA

F(2

011)

inth

eU

WR

atC

orv

alli

s,w

as44

4m

3/

sw

ith

MD

Fas

low

as15

0m

3/

s(U

SG

Sg

aug

ing

stat

ion

#14

1716

00).

TM

DL

s,th

eC

ity

of

Co

rval

lis

isco

nsi

der

ing

furt

her

trea

tmen

to

fth

eir

effl

uen

tin

aco

nst

ruct

edw

etla

nd

wit

hsu

bse

qu

ent

SE

Din

toth

eU

WR

via

the

HZ

.wIn

ast

ud

yco

mp

arin

gp

rop

osa

lsto

mee

tT

MD

Ls,

itw

asco

ncl

ud

edth

atad

dit

ion

altr

eat-

men

to

fm

icro

po

llu

tan

ts,

tem

per

atu

re,

and

nu

-tr

ien

tsw

ou

ldo

ccu

rw

ith

pas

sag

eth

rou

gh

the

HZ

,w

hic

hw

asq

ual

itat

ivel

yd

efin

edas

the

zon

eo

fac

tiv

esu

rfac

ew

ater

and

gro

un

dw

ater

ex-

chan

ge

inth

eri

ver

bed

.w

Th

eC

ity

of

Co

rval

lis’

SE

Dal

tern

ativ

eis

ap

ract

ical

exam

ple

of

ho

wn

atu

ral

HZ

pro

cess

esth

atat

ten

uat

eW

WD

Cca

nb

eu

sed

inm

anag

emen

t.If

imp

lem

ente

d,

this

exam

ple

wil

lse

rve

asa

mo

del

for

futu

reH

Zm

anag

emen

tp

roje

cts.

(con

tin

ued

)

484

Ta

bl

e1.

(Co

nt

in

ue

d)

Sit

eS

etti

ng

and

hy

dro

log

yP

ollu

tion

and

atte

nu

atio

nIm

pli

cati

ons

for

HZ

man

agem

ent

Pin

eR

iver

(PR

),A

ng

us,

On

tari

o,

Can

ada

Th

eP

Rd

rain

sa

348-

km

2b

asin

(70

cm/

yea

rm

ean

rain

fall

).x

A60

m-w

ide

PC

Eg

rou

nd

wat

erp

lum

eex

ten

ds

195

md

ow

ng

rad

ien

tfr

om

aD

NA

PL

sou

rce

of

PC

Elo

cate

db

enea

tha

dry

-cle

anin

gfa

cili

ty.x

Th

eM

AF

is2

m3/

s,an

dsu

mm

erb

ase-

flo

ws

can

dro

pto

1m

3/

s.x

,y

Th

en

ear-

stre

amzo

ne,

wh

ich

incl

ud

esth

eH

Z,

stro

ng

lym

od

ified

the

dis

trib

uti

on

,co

nce

ntr

atio

n,

and

com

po

siti

on

of

the

PC

Ep

lum

eb

efo

reb

ein

gd

isch

arg

edin

toth

eri

ver

.xE

xte

nsi

ve

anae

rob

icb

iod

egra

dat

ion

inth

eto

p2.

5m

of

the

stre

amb

edw

aso

bse

rved

,ev

enth

ou

gh

no

bio

deg

rad

atio

no

fth

eP

CE

plu

me

occ

urr

edin

the

up

gra

die

nt

aqu

ifer

.x

Infi

ltra

tio

no

fco

nta

min

ated

gro

un

dw

ater

into

stre

ams

isco

mm

on

inu

rban

area

s.z

Th

eP

Rsc

enar

iosu

gg

ests

that

HZ

man

agem

ent

can

red

uce

con

cen

trat

ion

so

fg

rou

nd

wat

erco

nta

mi-

nan

tsb

efo

reth

eyen

ter

into

riv

ers

inad

dit

ion

toin

crea

sin

gat

ten

uat

ion

of

surf

ace-

app

lied

WW

DC

.

Riv

erE

rpe

(RE

),G

erm

any

Th

eR

Ed

rain

sa

wat

ersh

edar

eao

f21

1k

m2

(55

cm/

yea

rp

reci

pit

atio

n).

Th

est

ream

bed

ism

ade

up

of

clo

gg

edan

dan

aero

bic

fin

e-sa

nd

yse

dim

ents

wit

hh

igh

con

ten

tso

fo

rgan

icca

rbo

nan

dn

utr

i-en

ts.*

,{It

rece

ives

effl

uen

tfr

om

sep

tic

tan

ksp

illw

ays

and

sev

eral

pri

vat

ean

dm

un

icip

alW

WT

Ps.

*T

he

larg

est

WW

TP

dis

char

ges

0.5

m3/

s,w

hic

hco

mp

rise

sas

mu

chas

81%

of

the

stre

amfl

ow

.*T

he

MA

Fis

1.2

m3/

sw

ith

MD

Fas

low

as0.

62m

3/

s.*,{

Mic

rop

oll

uta

nts

,in

clu

din

gcl

ofi

bri

cac

id,

nap

rox

en,

ket

op

rofe

n,

bez

afib

rate

,ib

up

rofe

n,

dic

lofe

nac

,an

din

do

met

hac

inh

ave

bee

nd

etec

ted

inth

eR

E,

and

con

cen

trat

ion

sd

ecre

ased

wit

hd

epth

inth

eH

Zto

ag

reat

erex

ten

tth

anco

uld

be

acco

un

ted

for

by

dil

uti

on

,su

gg

esti

ng

deg

rad

atio

n.*

Sam

-p

les

wer

eta

ken

toa

dep

tho

f1

mb

elo

wth

eri

ver

bed

,an

dth

ed

ilu

tio

nef

fect

was

qu

anti

fied

usi

ng

con

serv

ativ

etr

acer

s(b

ora

tean

dan

thro

-p

og

enic

gad

oli

niu

m).

*

Att

enu

atio

no

fW

WD

Cin

the

HZ

of

anef

flu

ent-

do

min

ated

,u

rban

stre

amw

assu

gg

este

db

yin

situ

mea

sure

men

tso

fd

egra

dat

ion

rate

sin

alo

wla

nd

stre

am.*

Fu

ture

stu

die

sth

atq

uan

tify

atte

nu

atio

no

fW

WD

Cin

the

HZ

wil

lal

soh

ave

toac

cou

nt

for

the

dil

uti

on

fact

or

fro

mth

em

ixin

go

fri

ver

wat

eran

dg

rou

nd

wat

er.

Ref

eren

ces

cite

din

the

tab

lear

ein

dic

ated

by

sup

ersc

rip

ts:

aL

auv

eran

dB

aker

,20

00;

bG

rim

met

al.,

2004

;c G

rim

met

al.,

2005

;dS

mit

h,

2000

;eT

rees

eet

al.,

2009

;f C

hen

etal

.,20

09;

gM

end

ezan

dB

elit

z,20

02;

hG

ross

etal

.,20

04;

i Din

get

al.,

1999

;j L

inet

al.,

2006

;kH

oeh

net

al.,

2007

;l P

lum

lee

etal

.,20

08;

mM

urp

hy

etal

.,20

00;

nV

erp

lan

cket

al.,

2005

;oB

arb

eret

al.,

2011

;pB

rad

ley

,20

08;

qD

enn

ehy

etal

.,19

98;

r Ro

sen

ber

ryet

al.,

2012

;s Ro

sen

ber

ryan

dP

itli

ck,2

009;

t McM

aho

net

al.,

1995

;uB

rad

ley

etal

.,20

09;v

Fer

nal

det

al.,

2001

;wK

enn

edy

/Je

nk

s,20

11;x

Co

nan

tet

al.,

2004

;yB

eeb

e,19

97;z

Ro

yan

dB

ick

erto

n,2

011;

*Lew

and

ow

ski

etal

.,20

11;{ P

CB

,20

01.

4-N

P,

4-n

on

ylp

hen

ol;

AP

EC

s,al

ky

lph

eno

xy

eth

oxy

late

carb

ox

yli

cac

ids;

DN

AP

L,

den

sen

on

aqu

eou

sp

has

eli

qu

id;

HZ

,h

yp

orh

eic

zon

e;M

AF

,m

ean

ann

ual

flo

w;

MD

F,

mea

nd

aily

flo

w;

NIW

TP

,N

og

ales

Inte

rnat

ion

alW

WT

P;

NT

A,

nit

rilo

acet

icac

id;

PC

E,

per

chlo

roet

hy

len

e;P

FC

,p

erfl

uo

roch

emic

al;

SE

D,

sub

surf

ace

effl

uen

td

isch

arg

e;T

MD

Ls,

tom

eet

tota

lm

axim

um

dai

lylo

ads;

WW

TP

,w

aste

wat

ertr

eatm

ent

pla

nt;

WW

DC

,w

aste

wat

erd

eriv

edco

nta

min

ant.

485

related strongly to redox conditions, pH, and temperature, iscritical to determine appropriate management practices forwater-quality managers.

Hyporheic flow enhances mass transfer of dissolved or-ganic compounds, nutrients, trace organic contaminants, andterminal electron acceptors [e.g., O2, manganese(IV), iron(III),and SO4

2 - ] to the subterranean habitats of both microorgan-isms and macroorganisms, thereby influencing transforma-tion rates (Harvey and Fuller, 1998; Bernhardt and Likens,2002; Gandy et al., 2007). HZ flow also results in contaminantscontacting subsurface sediments and the integuments ofaquatic biota, such as invertebrates and fish (Smith et al.,1998). Both transformation and sorption can be enhanced inthe HZ under the right conditions (Triska et al., 1993; Paria,2008).

Carbon. The HZ plays a paramount role in carbon cycling(Hendricks, 1993; Fellman et al., 2009). For example, riparianflora and fauna produce dissolved organic carbon (DOC),which supports biochemical processes that are vital to thehealth of streams (Fig. 1). DOC concentration in most unim-paired streams typically ranges up to 5 mg/L, while effluentfrom WWTPs typically contains between 5 and 30 mg/L (Imaiet al., 2002). Thus, organic matter from municipal effluent mayact as an additional source of DOC (Andersen et al., 2004).

Biofilms, that is, compilations of autotrophic and hetero-trophic microorganisms which coat the surfaces of both waterand substrate, in the HZ need DOC for their metabolism,causing the HZ to act as a DOC sink because of carbon uptakeand sequestration that occurs through resident bacteria andinvertebrates (Barlocher and Murdoch, 1989). For example,Findlay et al. (1993) found that the metabolic activity of hy-porheic organisms removed *50% of the DOC in interstitialwater in the HZ underneath a gravel bar of Wappinger Creek,an impacted urban stream in New York. Efficient water ex-change between the stream and the HZ accelerates the me-tabolism and increases consumption of DOC in the hyporheicwater. More information on biofilms is provided in the ecol-ogy section.

Several studies have shown that hyporheic ecosystems areheterotrophic (i.e., microbes use DOC as a C source in additionto fixing CO2 to sustain growth). Thus, the HZ serves as a linkbetween the river, which primarily consumes DOC, and thesurrounding ecosystem that provides the C. A comparisonbetween two small, pristine streams in Tennessee and NorthCarolina showed that relative to gross primary production(GPP), respiration was about 2.4 times greater in the streamwith the larger HZ; the value of GPP in the stream with thelarger HZ was 0.21 g O2/m2 per day compared with 0.32 inthe other stream (Mulholland et al., 1997). The authors con-cluded that metabolism was enhanced in the HZ comparedwith the stream. Heterotrophic conditions have also beenobserved in 4th- and 5th-Strahler stream order rivers in Mi-chigan (Uzarski et al., 2004), wherein an increase in nutrientand O2 supply associated with hyporheic flow was assumedresponsible for respiration.

In contrast, a much lower hyporheic respiration rate rela-tive to GPP was observed in the River Lahn in Germany, inwhich municipal wastewater effluent comprised 30% of baseflow (Ingendahl et al., 2009). At this site, the observed con-tribution of the HZ to riverine ecosystem respiration rates wasonly 14%, compared with HZ respiration rates from 40% to

96% observed in other studies (Naegeli and Uehlinger, 1997;Fellows et al., 2001; Battin et al., 2003). The respiration rates inthe upper 20 cm of sediment were approximately one order ofmagnitude greater than in sediments between 20 and 40 cmdepths. The authors suggested that the eutrophic state of theRiver Lahn enhanced algal primary production in the streamand decreased respiration in the HZ. Wagner and Beisser(2005) investigated the effect of addition of nutrients in a 2nd-Strahler stream order gravel stream, the Oberer Seebach inAustria, on biofilms and hyporheic biota. Both the quantityand quality (i.e., the ability to support diverse communities)of the biofilm increased during the 8 month experiment, andthe response of animals varied and was attributed to thequality of the biofilm; the living space for certain animalsincreased, and some exhibited increased reproductive suc-cess, while the mobility of others decreased.

Consequently, the composition and concentration of or-ganic C can influence the biological activity in the HZ.However, different sites can exhibit contrasting relationships.For example, biofilm growth in hyporheic sediments in StonyCreek in New Brunswick, Canada, was not related to DOCconcentration (Barlocher and Murdoch, 1989). In contrast,Findlay et al. (2003) observed a positive correlation betweenthe bacterial activity in biofilms and DOC concentration inWappinger Creek in New York. These conflicting findings arelikely related to whether C is a limiting nutrient in the eco-system.

The nature of organic carbon can also affect the ability ofthe microbial community to transform trace organic contam-inants, which are described in a later section. For example,microorganisms metabolizing on cattail and duckweed ex-tracts exhibited a better ability to remove certain recalcitrantpharmaceuticals in microcosms compared with microorgan-isms metabolizing wastewater effluent (Lim et al., 2008). Thehighest increase in degradation rate was observed for gemfi-brozil (a drug used to reduce lipid levels in blood), which wasnot degraded in DOC from wastewater effluent. However, itwas almost completely transformed after 6 days in plant ex-tracts, which is a residence time that could occur along deepflow paths in the HZ (e.g., residence times > 10 h were ob-served in a bedrock constrained reach of Lookout Creek inOregon (Kasahara and Wondzell, 2003). Differences betweenthe wastewater-derived DOC and the plant-derived DOC forthe existing communities were assumed responsible for thedifferent attenuation rates. Consideration of plants in additionto microbes may, thus, be important for optimizing contam-inant attenuation in the HZ.

Hydrologic modification can influence the concentrationand composition of DOC in streamwater and hyporheic wa-ter. For example, Westerhoff and Anning (2000) found thatstreams with a natural flow regime had higher DOC con-centrations compared with streams with regulated flow re-gimes characterized by damped temporal variability relativeto natural high and low flows. They also found that DOC instreams dominated by wastewater effluent was composed ofC molecules with lower molecular weight and less complexstructure, which provides poorer substrate for biofilms, andconsequently reduced food quality for the invertebrates andother organisms that feed on these biofilms. Although theorganic C in the HZ was not investigated, the hyporheic watermay be influenced by the stream water, especially alongdownwelling reaches with minimal groundwater exchange.

486 LAWRENCE ET AL.

Consequently, the diversity of the microorganisms and mac-roorganisms in the HZ of urban streams may be limited.

Nutrients. Excess nutrients are a major concern forstream-water quality and ecosystem health, and can result ineutrophication and anoxic conditions in the HZ. Nutrientbalance can be disturbed by discharges from WWTPs, returnflow from agricultural fields, and storm runoff (Withers andJarvie, 2008; Carey and Migliaccio, 2009). Both biotic andabiotic processes in the HZ attenuate N and P and mayprovide an especially important ecosystem service duringdry-season flows when urban streams are subject to thehighest concentrations of these nutrients (Withers and Jarvie,2008).

Denitrification occurs primarily by activities of the anaer-obic microbial community residing within the lower HZlayers (Fig. 1). It is controlled by relative rates in advectionand diffusion of N and O2 from the overlying stream waterinto the bed, with greater denitrification rates occurring underconditions of low oxygen flux (Claessens et al., 2010a, 2010b).Increased NO3

- levels in streams can increase denitrificationrates in sediments if residence times are long enough and thesediment zones are sufficiently voluminous, making the HZ asignificant nitrate sink (Cirmo and McDonnell, 1997). The HZmay develop anoxic conditions that promote denitrification inshallow, deoxygenated sediments of effluent-dominated ur-ban streams that are supplied by municipal wastewater ef-fluent and urban runoff (Schipper et al., 1993; Mayer et al.,2010; Lewandowski et al., 2011). In addition, groundwaterpassing through the HZ may provide significant NO3

- inputsto some surface waters, especially during base flow. Deni-trification rates in the HZ depend highly on the amount oflabile organic C available to serve as a microbial energysource, and, thus, N cycling is linked to C cycling (Holmeset al., 1996; Bernhardt and Likens, 2002; Strauss et al., 2002;Groffman et al., 2005; Birgand et al., 2007).

The efficiency of NO3- removal rates is highly variable and

depends on both the relative amount and the residence time ofstream water flowing into the HZ, and other variables, in-cluding seasonality, water temperature, pH, and riparianvegetation (Cirmo and McDonell, 1997; Groffman et al., 2005;Geza et al., 2010a; Zarnetske et al., 2011). The lowest NO3

-

concentrations typically occur in the summer months whenplant growth is greatest and there is relatively lower flow aswell as reduced nutrient loading (Birgand et al., 2007).Growing plants with their roots in the HZ take up N into theirtissues and thus act as a nutrient sink, but they can subse-quently release the N back into the water column on theirdeath or senescence (Dosskey et al., 2010). In addition to suchbotanic processes, heterotrophic immobilization or assimila-tion of N can decrease nitrification rates, while at the sametime consuming O2 and promoting denitrification (Mulhol-land et al., 2000; Baker et al., 2001; Groffman et al., 2005).

N transport in the HZ is closely linked to hydrologic flowpaths (Cirmo and McDonnell, 1997; Pinay et al., 2009). Thus,information on hydraulic conductivity, grain-size distribu-tion, streambed geometry, and stream velocity is needed toanalyze hyporheic N exchange (Salehin et al., 2004). Althoughlow O2 and increased NO3

- levels promote denitrification inthe HZ (Hill et al., 1998; Kasahara and Hill, 2006), the extent towhich denitrification occurs is determined principally byhydrology. High hydraulic conductivity not only promotes

large exchange volumes, but also limits residence time in theHZ, which is necessary for N removal (Grimaldi and Chaplot,1999). Surface-water storage capacity can also be a main de-termining factor of residence time in N attenuation zoneswithin the HZ (Hill et al., 1998), and denitrification rates arealso influenced by surface features, such as riffle steps, gravelbars, and meanders (Kasahara and Hill, 2006).

P activity is affected by a variety of conditions. For ex-ample, under reducing conditions commonly encountered inO2-depleted zones deep in the HZ of pristine streams, or atshallower depths in effluent-dominated urban streams, ironand manganese oxides dissolve and the metal ions migrateupward until conditions become oxic where they precipitateas hydroxides and oxides ( Jarvie et al., 2008). These freshlyprecipitated minerals, sometimes called the ‘‘oxic cap,’’ havelarge surface areas and can sequester P from the hyporheicwater within the HZ (Lijklema, 1980), thereby acting as abarrier for P release into the water column ( Jarvie et al., 2008).A combination of these mechanisms may be responsible forthe P attenuation observed downstream of a WWTP alongthe River Vene in France where 25% of the P in the effluentwas sequestered in the HZ (David et al., 2011). Conversely,release of P stored in hyporheic sediments of effluent-dom-inated urban streams has also been observed (Haggard et al.,2005).

A decrease in the thickness and the effectiveness of the oxiccap in attenuating PO4

3 - concentrations can result in degra-dation of organic matter in effluent, can consume O2 in hy-porheic sediments, and can cause metal oxides to reductivelydissolve at shallower depths ( Jarvie et al., 2008). For example,growth of sewage fungus caused reducing conditions and Prelease in shallow sediments in a small stream in easternEngland called the Lone Pine Pasture Stream (Palmer-Felgateet al., 2010); more than 30 mg/L P was observed in the HZ ofthis stream.

Biological uptake by bacteria and plants in the HZ has beenshown to remove or transform bioavailable PO4

3 - to lessavailable mineral forms. For example, the proportion of dis-solved to total P in both surface water and groundwater was‡ 90% in a chalk stream, the River Lambourn, in England,while in the HZ of this stream this proportion was < 70% as aresult of creation of particulate and colloidal forms of P(Lapworth et al., 2011). Similarly, greater PO4

3 - uptake wasobserved in a pristine 2nd-Strahler stream order with a largeHZ in North Carolina compared with a similar type streamwith a smaller HZ in Tennessee (Mulholland et al., 1997). Theuptake was, at least partially, related to biological activitywithin the HZ.

Trace organic contaminants. A wide range of these che-micals, which are synonymously referred to as emergingcontaminants, microcontaminants, trace pollutants, and mi-cropollutants, are derived from domestic and industrialwastewater and enter the urban water cycle via both WWTPsand storm water (Gobel et al., 2007; Ternes et al., 2008). Theyare typically defined to include minerals and chemicals thatare present in the environment in trace amounts, which have apotentially deleterious effect on public health or aquatic eco-systems, and they are not typically regulated with regard towater-quality criteria (Sedlak et al., 2000). Their concentrationsrange from ng to lg/L (Plumlee and Reinhard, 2007; Monteiroand Boxall, 2010). Common examples include pharmaceutical

ACTIVE MANAGEMENT OF HYPORHEIC ZONES IN URBAN STREAMS 487

compounds and household-cleaning products. These sub-stances can lead to toxicity in some aquatic biota despite beingpresent at very low concentrations, which presents a real butpoorly understood challenge to water management (Conn etal., 2010; Brar, 2011).

A few reports of unregulated trace organic contaminants inthe HZ are available (Hoehn et al., 2007), but a limited numberof studies report the attenuation of certain regulated, organicwastewater-derived pollutants in the HZ, which may behavein a similar manner, including fuel oxygenates (Landmeyeret al., 2010), tetrachloroethene or perchloroethylene (Conantet al., 2004), and volatile organic compounds (Durand et al.,2007). In general, the few published studies have found thatthe HZ has a high capacity for degradation of organic con-taminants, but that the volume of stream water entering theHZ is small relative to stream flow (Hoehn et al., 2007; Le-wandowski et al., 2011). Thus, long travel distances would beneeded to achieve sufficient residence time to provide water-quality enhancements.

Wastewater effluent is not the only source of trace organiccontaminants to streams and their HZs (Rayne and Forest,2009). For example, storm runoff from impermeable surfacesis a recognized contributor of such contaminants to urbanstreams (Oram et al., 2008; Zushi et al., 2008; Cousins et al.,2011; Grebel et al., 2013). Agricultural activities can releasepesticides, such as pyrethroids and organophosphates, tostreams, and these can flow into urban areas and have nega-tive impacts on biodiversity (Liess and von der Ohe, 2005;Martin, 2009). Nonetheless, wastewater effluent is the pri-mary source of most trace organic contaminants in urbanstreams. Clearly, more research in the area of organic con-taminant interactions with the HZ, particularly in urban oragricultural streams, is warranted.

Managers of water utilities are concerned about potentialrisks, including the possibility of contaminated streamwaterinfiltrating through the HZ to underlying aquifers and nega-tively impacting drinking-water sources. For example, thecommonly used wastewater-indicator compounds ethylene-diaminetetraacetic acid (EDTA) and naphthalene dicarboxylicacid (NDC) were found in drinking water production wellslocated at 1.8 km and 2.7 km downgradient from the SantaAna River in California and also in a monitoring well adjacentto the river (Ding et al., 1999) (Table 1); EDTA concentrationsmeasured in this study ranged from 0.1 to 6.3 lg/L, and NDCconcentrations ranged from undetectable to 1.5 lg/L. Incontrast, other commonly used indicators such as nitriloaceticacid and alkylphenoxy ethoxylate carboxylic acids were at-tenuated more rapidly in a soil-aquifer-treatment system andwere, thus, not encountered in the production wells in thisstudy. Concerns about aquifer contamination through the HZarising after the detection of perfluorochemicals prevented aplanned water reuse project in Coyote Creek, California (Ta-ble 1). However, some aquifers underlying urban areas arealready contaminated from legacy impacts, and downwellingof highly treated recycled water may actually improve water-quality conditions (Bischel et al., 2013).

Laboratory experiments have shown that biodegradationof trace organic contaminants occurs in hyporheic systems,and biodegradation may have been due to faster exchange,larger HZ area, longer residence time (e.g., due to deeper HZcaused by greater pressure head), or some combination ofparameters. Under aerobic conditions in a batch experiment

composed of water and sediment collected from the RoterMain, a river in Germany, biodegradation and sorption at-tenuated pharmaceuticals with half-lives ranging from 3 to328 days, suggesting that biodegradation in HZ sediments isan important pharmaceutical-attenuation mechanism (Loffleret al., 2005). In another study, laboratory flume experimentssimulating flow in this river found significant degradation ofbezafibrate, gemfibrozil, and ibuprofen when the exchange ofsurface water and pore water in the upper 20 cm of sedimentswas experimentally enhanced (Kunkel and Radke, 2008).Increases in the flow velocity in the flume created more oxi-dizing conditions and enhanced advective transport of con-taminants into the sediments, which accelerated degradation.Half-lives ranged from 1.2 to 6.9 days, and rate limitedtransfer was observed at low streamflow.

Apparently conflicting results obtained from laboratorystudies and field studies may be related to the difference ingeochemical conditions in the HZ. For example, in a follow-upfield study of pharmaceuticals in the Roter Main, only limitedattenuation was observed over a reach representing a traveltime of 6–30 h under anoxic conditions (Radke et al., 2010),which is a marked difference from the outcomes of the labo-ratory studies that were conducted. In situ measurements inthe field component of this study showed that photolysis wasof limited importance, and naproxen was the only pharma-ceutical observed to photodegrade with a half-life of *4 days.

Reactive tracers were used to assess attenuation of phar-maceuticals in the Sava Brook in central Sweden. Six com-pounds were studied, and only ibuprofen and clofibric acidwere attenuated with half-lives of < 0.5 day and 2.5 days,respectively (Kunkel and Radke, 2011). Due to the limited HZexchange through the fine-grained sediments in the stream,the authors hypothesized that the observed degradation re-sulted from biofilms growing on submerged macrophytesand sediment in the uppermost (*7 cm) HZ layer. In theRiver Erpe, which consists of *80% effluent, some attenua-tion of pharmaceuticals was observed in the top 100 cm of theHZ (Table 1), but the details of this attenuation were difficultto elucidate because of varying surface water composition.

The variability in attenuation rates of trace organic con-taminants observed among stream studies indicates largedifferences between streams in surface, hyporheic, andgroundwater conditions, illustrating the importance of site-specificity, and perhaps explaining some of the variability inthe concentrations of trace organic contaminants observedamong various aquatic reservoirs. The presence of O2 as anelectron acceptor appears to be very important for the bio-degradation of many trace organic contaminants, which oftendegrade aerobically (Meakins et al., 1994). Thus, degradationmay not be optimal in some effluent-dominated urbanstreams that have a relatively large O2 demand. The differentresults observed also highlight the lack of mechanistic un-derstanding of attenuation in both surface water and HZs.Aerobic and anaerobic degradation rates, and the rate ofsolute transport in the HZ, are poorly characterized.

Redox conditions. Oxygen depletion leading to localizedanoxia in the HZ occurs for multiple reasons, both abiotic andbiotic. These can include reductions in O2 concentrations in thesupplying stream water, reductions in infiltration through thestreambed because of reduced permeability (i.e., through sil-tation or channelization), switch from a downwelling to an

488 LAWRENCE ET AL.

upwelling condition as a result of changes in the groundwaterhead, increases in ambient temperatures that reduce O2 disso-lution, high levels of nutrients or biological O2 demand leadingto enhanced biological production and subsequent O2 deple-tion, and changes in the rates or pathways of microbe-mediatedchemical transformations because of alterations in environ-mental conditions (Heffernan et al., 2008; Nogaro et al., 2010). Ofparticular concern in urban streams are impermeable concreteflood-control channels that completely disconnect streams fromtheir HZ, thereby entirely eliminating the O2-supplying func-tion of stream water (Bernhardt and Palmer, 2007).

The redox conditions in the HZ represent a dynamic bal-ance between the influx of O2 as well as other electron ac-ceptors, and DOC, which consumes O2 as it undergoesbiodegradation. For example, turbulent flow aerates streamwater, resulting in O2 concentration near saturation, 8–10 mg/L, depending on water temperature. Stream waterDOC concentrations typically range from 1 to 20 mg/L(Worrall et al., 2004). In contrast, groundwater usually con-tains dissolved O2 concentrations around 1 mg/L, and DOC<1 mg/L (Malard and Hervant, 1999; Chapelle et al., 2012).The rate of organic C degradation depends on the chemicalstructure of the DOC, and highly aromatic organic matter ismore persistent (Thurman, 1985). Stream water is said to havea net positive O2 demand if the O2 demand from DOC exceedsthe dissolved O2 level. Continuous aeration of stream waterfrom turbulent mixing can keep O2 levels high, and turbulentmixing is enhanced by higher streamflows. Therefore, therelative mixing of the surface water and groundwater in theHZ largely determines the redox conditions, because a greatersurface water contribution typically results in more aerobicconditions.

The upper strata of hyporheic sediments is typically aerobicbecause of infiltration of aerated surface water, while bio-logical degradation consumes O2 and creates reducing con-ditions in the deeper sediments. Oxidation of DOC isthermodynamically most favorable with O2 as the primaryelectron acceptor followed by NO3

- , manganese oxides, iron(oxy)-hydroxides, and sulfate, in order of decreasing energyreleased during reaction. When O2 in the sediments is con-sumed, NO3

- is converted to N2 gas, followed by reductivedissolution of manganese and iron oxides and consumption ofsulfate. In pristine streams, the oxygenated area of the HZtypically extends to greater depths than in lowland streams,which exhibit less turbulence and aeration, and typically havefiner grained sediments (i.e., sands, silts, and clays) that aremore prone to clogging and thus less hyporheic exchange(Findlay, 1995; Boulton et al., 2011). Hill et al. (1998) observedthat O2 concentrations declined with depth within the HZ ofan agricultural stream in Ontario, Canada; the concentrationsmeasured were greatest at the head of riffles where streamwater infiltrated. Similarly, other researchers have observedchanges in O2 concentration in the HZ caused by down-welling stream water (Hancock et al., 2005). In anotherlowland stream that receives a high wastewater supply, O2

consumption within the upper millimeters of the stream-bed sediments created reducing conditions in the HZ (Le-wandowski et al., 2011).

Redox conditions vary significantly in the HZ, and stronglyaffect biological activity. Given the appropriate redox condi-tions for a particular contaminant, efficiency of biodegrada-tion in the HZ is enhanced by microbes and potentially

facilitated by burrowing animals, such as benthic macro-inverterbrates. Biodegradation rates can be much faster forsome pharmaceuticals under aerobic than under anaerobicconditions (Groning et al., 2007; Kujawa-Roeleveld et al., 2008;Musson et al., 2010). In laboratory experiments, faster atten-uation rates of selected pharmaceuticals that degrade aero-bically have been observed in the upper part of the HZ(Kunkel and Radke, 2008). Respiratory and hydrolytic activityof bacteria is strongly linked to O2 consumption, and thisbacterial activity is generally greater within oxygenated sed-iments, which has been demonstrated in slow-filtration col-umn experiments (Mermillod-Blondin et al., 2005). In a studyof the White Clay Creek in Pennsylvania, hyporheic respira-tion accounted for 41% of whole ecosystem respiration withmost of the biological activity documented in the upper 30 cmof the sediments in reaches where HZ exchange was observed(Battin et al., 2003). Burrowing animals in the HZ createpreferential flow channels that can enhance infiltration,thereby extending the aerobic zone to deeper sediments(Boulton et al., 1998).

Generally, redox conditions in the HZ change along astream reach, and stream water will access both aerobic andanaerobic compartments. Thus, favorable conditions for bio-degradation can exist for any contaminant provided residencetimes in the appropriate redox zones are sufficient. However,most trace organic contaminants degrade more efficiently inaerobic conditions (Meakins et al., 1994; Angelidaki et al.,2000), which tend to occur in the HZ during higher streamflows or under stream channels with surface features such aslarge boulders and large woody debris that promote turbulentmixing. During high flows, the volume percent of streamflowthrough the HZ can be small, which would limit attenuation.Thus, reliable removal of contaminants is more likely instreams in which both the majority of flow occurs through theHZ and the flow is aerated, which may be limited to specificmorphologies (e.g., course grained substrates, riffle-poolsequences, channel meanders, and other complex surfaceelements).

Metal concentrations in the HZ are also primarily con-trolled by redox conditions. For example, under reducingconditions, manganese oxides and iron (oxy)hydroxides dis-solve, releasing manganese and iron to the stream; however,manganese and iron oxides have a strong affinity to bindother metals, and, thus, they can be co-transported (Dzombakand Morell, 1990; Harvey and Fuller, 1998; Smedley andKinniburgh, 2002). In contrast, infiltration of aerated surfacewater causes dissolved manganese and iron to oxidize andprecipitate as oxide and hydroxide minerals with other metalsco-precipitating or sorbing to freshly oxidized surfaces.

The HZ may prevent groundwater contamination becauseof its redox chemistry. For example, in mining impacted areas,the HZ has provided a barrier to metals (Gandy et al., 2007).At the Silver Bow Creek in western Montana, acidic, metal-contaminated groundwater mixed with stream water andiron formed oxides in the HZ to a depth of 80 cm, whilemanganese precipitated in the shallower, more highly oxi-dized sediments (Benner et al., 1995). In this study, Benneret al. (1995) suggested that the geochemical boundary of ironcorresponded to the interface between the groundwater andthe deeper HZ, while the geochemical boundary of manga-nese corresponded to the boundary between the water col-umn and the shallower HZ.

ACTIVE MANAGEMENT OF HYPORHEIC ZONES IN URBAN STREAMS 489

Acidity. The acidity in the HZ, which is typically re-presented as the pH, is a relatively easy-to-measure mastervariable that affects chemical and biological processes. Forexample, metal sorption to sediment tends to decrease withpH (Stumm and Morgan, 1996). The pH of the HZ is a func-tion of the relative acidities of the contributing groundwaterand surface water, and it is also correlated to other parame-ters, such as O2 and DOC. Groundwater is typically wellbuffered near neutral pH, although mining impactedgroundwater can be more acidic, and, therefore, the pH of thehyporheic water in mining areas may be lower (Gandy et al.,2007).

Oxidative precipitation of metal hydroxides consumes O2

and hydroxide ions, thus lowering the pH. In the Pinal CreekBasin in Arizona, for example, manganese-contaminated,acidic groundwater continually mixed with oxygenatedneutral stream water, removing 20% of the manganese load inthe HZ (Harvey and Fuller, 1998). The increase in O2 and pHwithin the HZ of this system controlled the microbially en-hanced process of manganese oxidation. In contrast, reductivedissolution of metal oxides can increase pH. Petrunic et al.(2005) found that pH increased as microbially mediated re-ductive dissolution of manganese oxides occurred in columnexperiments that were designed to simulate conditions ofaquifer recharge by rivers.

The pH can also be lowered in the HZ by production oforganic acids by microbes during their processing of organicmatter. For example, in a temperate forested stream innorthern Wisconsin that receives large seasonal inputs of leaflitter, the pH decreased from 8.5 to below 7.0 within the HZ(Schindler and Krabbenhoft, 1998). This drop in pH wascorrelated to increases in methane and DOC, and to the bio-chemical activity of microbes processing the leaves.

Temperature. Water temperature is an important deter-minant of environmental conditions in the HZ because of itsdirect correlation with both O2 levels and biochemical reac-tion rates and pathways (Hatch et al., 2006; Schmidt et al.,2006; Webb et al., 2008). Emerging technologies for high-res-olution measurement of water temperatures in the HZ includedistributed sensors that rely on fiber optics (Briggs et al., 2012).The temperature of hyporheic water reflects the relative con-tributions of both surface water and groundwater (Krause etal., 2009; Rau et al., 2010). Surface-water temperatures followambient air temperatures with a relatively short time lag, andthe temperature of smaller streams is more responsive to localatmospheric microclimates than the temperature of largerrivers because of their lower thermal mass (Nelson and Pal-mer, 2007). In contrast, groundwater temperatures remainrelatively constant near the mean annual air temperature;however, smaller perched groundwater reservoirs may beinfluenced by stream temperature fluctuations (Chu et al.,2008).

The temperature regime of the HZ is, thus, responsive tohuman alterations to stream-water temperatures, which oc-cur through a variety of mechanisms at different spatio-temporal scales, including WWTP discharges, dam releases,and climate change (Kaushal et al., 2010). Wastewater efflu-ents are typically warm relative to natural rainfall, snow-melt, or groundwater and can, thus, raise streamtemperatures (Kinouchi et al., 2007). The temperature ofwater released from dams depends largely on whether it is

released from the bottom or the top of the impounded res-ervoir (Olden and Naiman, 2010). Climate change will likelyincrease regional streamwater and, consequently, hyporheicwater temperatures, with consequent effects on ecologicalcommunities (Lawrence et al., 2010). Further, climate-induced changes in the western United States may have ledto widespread forest infestations that influence the temper-ature, hydrology, and ecology of these urban corridors(Mikkelson et al., 2013).

Ecology and the HZ

The HZ is an important habitat for many freshwater or-ganisms, which are often referred to as the hyporheos whenthey make use of this subterranean zone (Boulton et al., 1998;Wood et al., 2010). Some of these organisms only use the HZtransiently as a refuge when the conditions on the surfacebecome unfavorable, such as during floods, droughts, or pe-riods of impaired water quality or high predation pressure(Stubbington et al., 2011). Others use the HZ exclusively, andthese have been observed to exhibit a higher level of ende-mism relative to biota in surface-water habitats because of therelatively lower level of disturbance (Gibert and Deharveng,2002). In some cases, the ecological connection between thesurface water and the HZ has been found to be very weak (i.e.,the aquatic organisms do not appear to migrate between thesetwo reservoirs), but often this is not the case. The reason forthis lack of connectivity is not well understood, but it may berelated to the highly variable nature of subsurface flow pathsand also to the unique adaptations of biota, which in manylocations are highly site specific (Dole-Olivier, 2011; Younget al., 2011).

The HZ is also an important component of the life cycle ofmany of these biota, including fishes, macroinvertebrates, andamphibians (Lopez-Rodrıguez et al., 2009; Resh and Rosen-berg, 2010; Williams et al., 2010). For example, many salmonidfishes deposit their eggs in the HZ, and egg survival dependson localized downwelling of streamwater to both meet oxy-gen requirements and provide nutrients during early matu-ration (Soulsby et al., 2009). Downwelling may expose eggs tochemical contaminants in wastewater-effluent-impactedstreams, which is a significant ecological concern. Manyaquatic insects use the HZ during their young larval stagesand before emerging from both the HZ and the stream tooccupy terrestrial habitats during their adult life stages(Reynolds and Benke, 2012). Amphibians, such as salaman-ders, may use the HZ to forage, escape predation, obtainrefuge from low streamflows, and also for nesting (Feral et al.,2005).

To survive in the HZ, biota must have specialized biolog-ical traits, including the ability to burrow through the sub-strate, navigate in the absence of light, and obtain sustenancefrom the relatively limited food sources available, such asmicrobes and partially decayed organic matter (Bonada et al.,2007; Mueller et al., 2011; Datry, 2012). For example, macro-invertebrates with burrowing ability often have relativelystrong forearms to dig through sediments, and those thatspend most of their time in the HZ often have reduced eyes orlack of eyes altogether as a result of the light limitation(Walters, 2011). Small body size is another common trait forhyporheic organisms as an adaptation to both the low flowand the low food availability (Boulton, 2007).

490 LAWRENCE ET AL.

The hyporheic fauna are sampled using a variety of de-vices, and the device selected affects the densities measuredand also the proportion and size of the taxa observed. Forexample, hyporheic sampling devices can include nonfrozenor frozen sediment corers, colonization chambers filled withartificial substrates, and pumps or bailers attached to stand-pipes or wells (Fraser and Williams, 1997). Since the numbersof organisms are often extrapolated and then expressed in cm3

or m3, the densities reported can be very high. Microorgan-isms are also an important component.

Biofilms serve an important function in water-quality en-hancement through natural processes (Cardinale, 2011; Hsuet al., 2011; Writer et al., 2011). They have a three-dimensionalmatrix that supports interconnected communities of algae,bacteria, and fungi (Lear and Lewis, 2009; Augspurger et al.,2010; Pohlon et al., 2010). The maintenance of the biofilm isperformed naturally by grazers, typically macroinvertebrates,which keep the constituent communities in an active growthstate that provides optimal conditions for processing bothorganic and inorganic compounds as well as other chemicalconstituents in the aqueous environment, such as trace metals(Sabater et al., 2002).

The important role of biofilms in WWTPs and in con-structed wetland systems is widely recognized (McBride andTanner, 1999; Toet et al., 2003; van den Akker et al., 2011;Jasper et al., 2013), but their role in regulating processes in theHZ is not nearly as well understood. For example, the com-munities comprising biofilms are responsible for many of theimportant redox reactions that result in the transformation oftrace organic contaminants (Gadd and White, 1993; Hockinand Gadd, 2003). Similar processes likely occur in the HZ;however, the community composition of biofilms in the HZmay be different than in surface waters.

Oxygen depletion can have dramatically negative effectson the life cycles of the hyporheos (Malard et al., 2002; Sabaterand Tockner, 2010; Tomlinson and Boulton, 2010). For ex-ample, salmonid eggs, which are buried in the HZ in struc-tures commonly known as redds, will fail to develop properlyin the absence of O2, resulting in reduced fecundity of salmonpopulations (Tonina and Buffington, 2009). In addition,macroinvertebrates in the HZ are highly sensitive to O2 in theinterstitial water, and their populations can be dramaticallyreduced or even completely extirpated if the local O2 supply ishighly curtailed or eliminated (Olsen et al., 2010).

Excessive deposition of fine sediment is a byproduct ofhuman development and has had dramatic negative effects onthe hyporheos, as well as on the surface-water biota, of manyurban streams (Paul and Meyer, 2001; Coleman et al., 2011).Increased siltation is often associated with conversion of for-est lands to other uses or with construction of roads andbuildings in urban areas. Fishes, macroinverterbrates, andamphibians have been observed to be adversely affected bysiltation effects (Yamada and Nakamura, 2009; Larsen andOrmerod, 2010; Sternecker and Geist, 2010; Sullivan andWatzin, 2010; Louhi et al., 2011). Increased sediment instreams from urban development may also lead to increasedP levels in stream waters (Geza et al., 2010b).

The deleterious effects of siltation result largely from theclogging of interstitial spaces between sand grains, gravels,and cobbles on the streambed (Arnon et al., 2010; Song et al.,2010). With pore spaces clogged, streamwater is no longerable to downwell into the HZ, which can cause the HZ to

become water stressed (Kasahara et al., 2009). In addition, finesediment can coat the gill surfaces of organisms, reducing O2-uptake efficiency and compounding the already deleteriouseffects of the associated reductions in O2 delivery (Kemp et al.,2011). However, many invertebrates and other organisms canrestore permeability in the streambed through the process ofbioturbation, which describes the formation of preferentialflow paths and the mixing of sediment that occurs as a resultof burrowing activities (Nogaro et al., 2010).

Many freshwater biota have physiologically optimal tem-perature ranges and if the local conditions are outside of thisrange, then the biota will be stressed or perhaps even elimi-nated (Li et al., 2011). Moreover, if water temperatures areoutside of the range of natural variability for a sufficientlylong time, nonnative species can overtake native species andreductions in biodiversity can occur (McDermott et al., 2010).Consequently, Olden et al. (2006) argue that watershed man-agers should aim at restoring not just natural flow regimes,but natural thermal regimes as well.

Active management of the HZ

Effluent-dominated, low-flow urban streams. These streamspresent a great opportunity to evaluate the efficacy of the HZto improve water quality or serve as a barrier against aquifercontamination. Streams during low-flow conditions, whichcan have long residence times and a high percentage of thetotal streamflow passing through the HZ, appear more likelyto exhibit measureable attenuation via sorption and biodeg-radation, provided redox conditions are suitable (e.g., suffi-cient O2 is available for trace organic contaminants thatundergo aerobic degradation).

Consequently, we believe that there is promise for water-quality enhancements for effluent dominated streams in aridenvironments, particularly during the late summer, fall, andearly winter, when streamflow is relatively low and HZ flowsmay be more significant. For example, recycled water addedto dry urban streambeds, which are often dry because of an-thropogenic activities such as ground-water pumping, mayenable revitalization of the ecology of these areas while en-hancing water quality of urban streams. We also believe thatengineering HZs in reconstructed urban streams may havepromise for providing water-quality enhancement.

Examples from the literature of low-flow, effluent-dominated urban streams with documented pollutant atten-uation illustrate implications for active management in a va-riety of geographic settings and under a range of scenariosrepresenting different HZ characteristics (Table 1). In theseurban basins, mean annual rainfall ranged from 33 to 114 cm,watershed areas varied from 48 to 63,000 km2, and mean an-nual streamflows were measured between 0.084 and 444 m3/s(Table 1). The measurements collected at these sites includedextensive hydrogeologic surveys and/or intensive analyses ofnutrients and trace organic contaminants.

Stream restoration. Restoration projects in streams areincreasingly being implemented to counteract the escalatingnegative effects of urbanization (Kondolf and Micheli, 1995;Ebersole et al., 1997; Palmer et al., 2009), but they typicallytarget surface features of streams without directly consid-ering the effects on the underlying HZ in the project designs(Hester and Gooseff, 2010). These projects often includechannel reconfigurations (i.e., installing meanders),

ACTIVE MANAGEMENT OF HYPORHEIC ZONES IN URBAN STREAMS 491

establishment of riffle-pool sequences, placement of struc-tures to increase physical-habitat complexity (i.e., installingboulders and large wood), selective planting of riparianvegetation, re-stocking of imperiled aquatic species, gravelreplenishment, and gravel cleaning operations (Sarriquetet al., 2007; Meyer et al., 2008; Bernhardt and Palmer, 2011).Gravel cleaning operations in the form of managed high-flow releases are intended to cleanse the bed of fine sedimentand may be especially relevant to WWTP-effluent projects assuggested by the example of the Santa Cruz River in Arizona(Table 1). Some of these practices may also have serendipi-tous benefits to contaminant removal because of the aerationfunction they provide.

Many of these restoration techniques that are applied at thesurface can directly affect the HZ. For example, installation ofboulders and large wood will promote localized increases inthe water-surface elevation (i.e., will effectively create smallwater-storage reservoirs), which will cause greater infiltration(potentially of oxygenated water) into the HZ provided thestreambed is sufficiently permeable (Lautz and Fanelli, 2008;Scordo and Moore, 2009; Boulton et al., 2010). Many types ofriparian vegetation will also enhance infiltration of water intothe HZ, because their roots create macropores that can act assubsurface flow paths (Dosskey et al., 2010), but these sameplants can also withdraw large volumes of water throughevapotranspiration and thereby reduce the total volume of thehyporheic reservoir (Pollen-Bankhead and Simon, 2010).However, evapotransporation may also potentially enhancedownwelling, because water removed by plants may be re-placed by stream water.

Benthic macroinvetebrates are also recognized as playingan important bioturbation role in the HZ and can, thus, en-hance permeability. They can break down particulate organicmatter into smaller pieces through their food-collection ac-tivities (which can be then be more easily consumed by mi-crobes) (Collins et al., 2007), and they can serve as a usefulindicator for the success of ecosystem recovery and restora-tion (Del Rosario and Resh, 2000). Numerous structural andfunctional metrics are available for determining the biologicalintegrity of the ecosystem based on macroinvertebrate com-munity composition (Sponseller et al., 2010; Wesener et al.,2011). Riverine wetlands and the lateral components of theHZ into the floodplain are important sources and sinks formany macroinvertebrates (Paillex et al., 2009).

Nutrients and trace organic contaminants are a concern forhuman health and biodiversity that should be considered instream restoration, and the HZ can play an important role(Higgins et al., 2010; Williams et al., 2010; Biksey et al., 2011;Lewandowski et al., 2011). For example, contaminated streamwater can pass through the HZ to groundwater aquifers,which are often used for drinking water (Krause et al., 2011).In contrast, contaminated water can also travel fromgroundwater to streams, and the example of the Pine River inOntario, Canada, where extensive anaerobic biodegradationwas observed as occurring in the top 2.5 m of the streambed,suggests that the HZ can provide an effective layer of pro-tection (Table 1). The management outcomes could be par-ticularly relevant to urban streams located near leaking(exfiltrating) underground sewers, which are common inmany older cities. Public health concerns exist related to traceorganic contaminants (although the risk has been found to bevery low in most cases examined), and there is some evidence

of potential harm to fishes and other freshwater biota that areexposed to such contaminants (Glassmeyer et al., 2005; Vajdaet al., 2008). In Boulder Creek, Colorado (Table 1), antide-pressants were measured in native white suckers (Catostomuscommersoni) and in HZ sediments (Schultz et al., 2010). Thiscreek also serves as a heavily used recreational area for thelocal human community, and public health is thus highlyrelevant.

Management opportunities. A variety of HZ manage-ment techniques for water reuse for stream restoration areenvisioned that require a more thorough understanding of theinteraction between the hydrology, chemistry, and ecology ofthe HZ in urban-stream systems. The stream is not just a pipeconveying water to a larger pipe, as is sometimes conceptu-alized by wastewater managers (Bencala et al., 2011). It hasboth ecological and aesthetic value. In water-stressed streamsand their associated HZs, reclaimed water can serve as awater supply for flow augmentation to provide furthertreatment or restore ecological condition (Fig. 1), particularlygiven the unlikely availability of alternative supplies of sur-face water or groundwater in the future.

The reclaimed water could be percolated throughstreambeds, stream banks, or floodplains (Fig. 1), dependingon the site-specific conditions, to achieve objectives that areimportant to local stakeholders. Surface features, such asriffle-pool sequences and large-wood installations, could beengineered to increase residence time. If sufficient energy isavailable at a low cost, water could be pumped from shallowwells downstream and injected back into the system up-stream to achieve greater residence time over a shorterstream reach. The example of the Willamette River in Cor-vallis, Oregon, suggests that subsurface effluent dischargeinto the HZ, preceded by treatment in constructed wetlands,can provide additional treatment and thus an additionallayer of protection before flows reach the river (Table 1).Engineered elements could be incorporated in the streamchannel to enhance aeration, and effluent discharges couldbe tailored to match ecologically significant components ofthe water-quality and flow regimes. Lastly, the relationshipof the HZ with physical channel modifications, such as checkdams that raise ground water levels, reduce channel erosion,and reduce local stream gradient, as well as the removal orreplacement of impermeable channel linings, should beconsidered. Such physical modifications are already beingimplemented for ecological, recreational, and other purposes(e.g., Strawberry Creek in Berkeley, California) (Charbon-neau and Resh, 1992).

As a potential engineered natural system to complementWWTPs, water could be routed from a treatment facility to astream channel where the substrate composition of HZs couldbe designed as originally suggested by Vaux (1968) and re-cently revisited by Ward et al. (2011), possibly using recycledconstruction debris or gravel imported from an externalsource. Alternatively, the water could be delivered to an un-derutilized channel or a paleochannel (i.e., an ancient, cur-rently inactive stream channel) where the substratecharacteristics are amenable to water-quality enhancement. Inaddition, carbon filters with higher hydraulic conductivitythan the surrounding substrate could potentially be added toHZs to enhance treatment and could be designed so that theycould be removed periodically for cleaning or replacement,

492 LAWRENCE ET AL.

although this technique has not been tested. Manipulationsthat promote favorable HZ redox conditions should also beconsidered in addition to these structural applications. Wepropose that engineering enhanced contaminant removal inartificial or modified HZs is analogous to the accepted prac-tice of reclaimed water treatment in constructed wetlands,providing both water-quality improvement and habitat en-hancement.

Barriers and research or logistical needs. To optimizeimplementation of active HZ management strategies, fur-ther knowledge is needed of how the associated flow aug-mentations or substrate modifications might affect drinking-water reservoirs and the diverse ecological communities inthe subsurface (fungal, microbial, floral, vertebrate, and in-vertebrate). The ecological communities in the HZ have notreceived as much attention scientifically as their counter-parts on the surface, but they deserve equal attention inbiodiversity conservation efforts. This idea should be in-troduced into the public consciousness. These ecologicalcommunities certainly play a paramount role in the regula-tion and maintenance of hydrologic flow paths, chemicalreaction rates, and contaminant fate in the subsurface HZenvironment.

The concept of active HZ management has been applied inonly a very limited number of cases in urban areas, and datacollection from these cases was often minimal, thus limitingour ability to evaluate the success of these projects (Table 1). Itappears that the focus of research in the near term should beon effluent-dominated, low-flow streams in which attenua-tion of nutrients and degradation of trace contaminants islikely to be reliably high. A related issue is that metrics formonitoring conditions in the HZ are not well established, and,thus, appropriate metrics of both water quality and ecologyshould be developed. Moreover, research will need to addresslegitimate concerns that exist over the potential for bothstreamwater and groundwater contamination which couldoccur if HZs are intentionally made more permeable or filledto higher capacity. In addition to the contaminants described,the HZ is also involved in trafficking of pathogenic bacteria,which could have public-health implications (Grant et al.,2011).

New laboratory experiments and field studies should in-form future model development. A wide variety of modelshave been adapted and applied to the HZ (Runkel, 1998;Lautz and Siegel, 2006; Kasahara and Hill, 2008; Claessensand Tague, 2009; O’Connor et al., 2010; Ward et al., 2011), butthese models need further development and testing to bewidely applied in active management scenarios. Develop-ment and refinement of such models requires tracer studiesto calibrate residence times, and preliminary studies haveoccurred in urban streams (Ge and Boufadel, 2006; Ryan andBoufadel, 2006a, 2006b; Ryan and Boufadel, 2007; Ryan et al.,2010; Ryan et al., 2011; Toran et al., 2012). Conservativetracers are often used to visualize breakthrough curves,which are analyzed using the reactive-solute-transport andtransient-storage concepts (Choi et al., 2000; Briggs et al.,2009; O’Connor et al., 2010). However, some subsurface-flowpaths may be too long to be detected by such tracer experi-ments (Bencala et al., 2011), and turbulence through inter-stitial spaces is difficult to capture mathematically (Grantand Marusic, 2011).

Moreover, increased model sophistication has not gener-ally led to greater reliability (Wondzell et al., 2009), andmodels based entirely on geomorphic data (as opposed todata from tracer experiments) may provide reasonable esti-mates of residence times in the HZ (Cardenas et al., 2004;Cardenas and Wilson, 2007; Cardenas, 2008). However, state-of-the-art models for the HZ have not been sufficiently linkedmechanistically with hydrochemical processes, bioturbationfrom macrofauna, such as fishes and macroinvertebrates, orvegetation uptake-processes. These variables can clearly in-fluence much of the variability in the biogeochemistry of theHZ, and, subsequently, pollutant attenuation, which as pre-viously discussed is more likely in streams in which both themajority of flow occurs through the HZ and the flow isaerated.

Conclusion

Under appropriate conditions, HZ management offers agreat and largely unexamined opportunity for improvingwater quality and supporting biodiversity in a rapidly ur-banizing world that is facing the uncertainty of global climatechange and continually changing cultural institutions andvalues. The HZ will inevitably respond to wastewater dis-charges whether water managers are aware of it or not, andincreased awareness will enable these natural processes to bemanaged more holistically. This awareness issue for watermanagers is illustrated clearly by the example of the SouthPlatte River in Colorado in which water utilities were un-knowingly affecting hyporheic exchange processes withhourly fluctuations of their effluent discharge rates (Table 1).We envision that water utilities of the future will include HZmanagement of effluent-dominated streams in their portfolioof projects intended to provide public value, and that waterwill flow out of cities cleaner than it flows in, and in a formwhich will benefit the ecology. However, barriers should beovercome before this vision can become a reality, and prog-ress will most likely be made by engaging multiple stake-holders. These stakeholders should include the general publicwho are the rate-payers, as well as the water professionalsworking in affiliated utilities, nonprofits, government, andacademic institutions.

Acknowledgments

This work was made possible through funding providedby the National Science Foundation Engineering ResearchCenter for Re-inventing the Nation’s Urban Water Infra-structure, ReNUWIt (NSF EEC-028968). The authors thank M.Plumlee for reviewing all or portions of this article, and K.Bencala and J. Zarnetske for providing resources andthoughtful discussion.

Author Disclosure Statement

No competing financial interest exists.

References

Andersen, C.B., Lewis, G.P., and Sargent, K.A. (2004). Influenceof wastewater treatment effluent on concentrations and fluxesof solutes in the Bush River, South Carolina, during extremedrought conditions. Environ. Geol. 11, 28.

ACTIVE MANAGEMENT OF HYPORHEIC ZONES IN URBAN STREAMS 493

Angelidaki, I., Mogensen, A.S., and Ahring, B.K. (2000). De-gradation of organic contaminants found in organic waste.Biodegradation 11, 377.

Arnon, S., Marx, L.P., Searcy, K.E., and Packman, A.I. (2010).Effects of overlying velocity, particle size, and biofilm growthon stream–subsurface exchange of particles. Hydrol. Process.24, 108.

Augspurger, C., Karwautz, C., Mußmann, M., Daims, H., andBattin, T.J. (2010). Drivers of bacterial colonization patterns instream biofilms. FEMS Microbiol. Ecol. 72, 47.

Baker, L.A., Hope, D., Xu, Y., Edmonds, J., and Lauver, L. (2001).Nitrogen balance for the central Arizona–Phoenix (CAP) eco-system. Ecosystems 4, 582.

Barber, L.B., Antweiler, R.C., Flynn, J.L., Keefe, S.H., Kolpin,D.W., Roth, D.A., Schnoebelen, S.J., Taylor, H.E., and Ver-planck, P.L. (2011). Lagrangian mass-flow investigations ofinorganic contaminants in wastewater-impacted streams. En-viron. Sci. Technol. 45, 2575.

Barlocher, F., and Murdoch, J.H. (1989). Hyporheic biofilms—Apotential food source for interstitial animals. Hydrobiologia184, 61.

Battin, T.J., Kaplan, L.A., Newbold, D., and Hendricks, S.P.(2003). A mixing model analysis of stream solute dynamicsand the contribution of a hyporheic zone to ecosystem func-tion. Freshw. Biol. 48, 995.

Beebe, J.T. (1997). Fluid Patterns, Sediment Pathways and WoodyObstructions in the Pine River, Angus, Ontario [PhD disserta-tion]. Waterloo, Canada: Department of Geography and En-vironmental Studies, Wilfrid Laurier University.

Bencala, K.E., Gooseff, M.N., and Kimball, B.A. (2011). Re-thinking hyporheic flow and transient storage to advanceunderstanding of stream-catchment connections. Water Resour.Res. 47, W00H03.

Benner, S.G., Smart, E.W., and Moore, J.N. (1995). Metal be-havior during surface-groundwater interaction, Silver BowCreek, Montana. Environ. Sci. Technol. 29, 1789.

Bernhardt, E.S., and Likens, G.E. (2002). Dissolved organic car-bon enrichment alters nitrogen dynamics in a forest stream.Ecology 83, 1689.

Bernhardt, E.S., and Palmer, M.A. (2007). Restoring streams inan urbanzing world. Freshw. Biol. 52, 738.

Bernhardt, E.S., and Palmer, M.A. (2011). Evaluating river res-toration. Ecol. Appl. 21, 1925.

Bianchin, M.S., Smith, L., and Beckie, R.D. (2011). Defining thehyporheic zone in a large tidally influenced river. J. Hydrol.406, 16.

Biksey, T.M., and Gross, E.D. (2001). The hyporheic zone: linkinggroundwater and surface water—Understanding the para-digm. Remediat. J. 12, 55.

Biksey, T.M., Schultz, A.C., Bernhardt, A.M., Marion, B., Pe-terson, C., and Smith, P. (2011). Ecological and human healthrisk assessment. Water Environ. Res. 2011, 1876.

Birgand, F., Skaggs, R.W., Chescheir, G.M., and Gilliam, J.W.(2007). Nitrogen removal in streams of agricultural catchments–a literature review. Crit. Rev. Environ. Sci. Technol. 37, 381.

Bischel, H.N., Lawrence, J.E., Halaburka, B.J., Plumlee, M.H.,Bawazir, S., King, J.P., McCray, J.E., Resh, V.H., and Luthy,R.G. (2013). Renewing urban streams with recycled waterfor streamflow augmentation: Hydrologic, water quality,and ecosystem services management. Environ. Eng. Sci. 30,455.

Boano, F., Demaria, A., Revelli, R., and Ridolfi, L. (2010a). Bio-geochemical zonation due to intrameander hyporheic flow.Water Resour. Res. 46, W02511.

Boano, F., Revelli, R., and Ridolfi, L. (2010b). Effect of stream-flow stochasticity on bedform-driven hyporheic exchange.Adv. Water Resour. 33, 1367.

Bonada, N., Rieradevall, M., and Prat, N. (2007). Macro-invertebrate community structure and biological traits relatedto flow permanence in a Mediterranean river network. Hy-drobiologia 589, 91.

Boulton, A.J. (2007). Hyporheic rehabilitation in rivers: Restoringvertical connectivity. Freshw. Biol. 52, 632.

Boulton, A.J., Datry, T., Kasahara, T., Mutz, M., and Stanford,J.A. (2010). Ecology and management of the hyporheic zone:Stream–groundwater interactions of running waters and theirfloodplains. J. N. Am. Benthol. Sci. 29, 26.

Boulton, A.J., Findlay, S., Marmonier, P., Stanley, E.H., and Va-lett, H.M. (1998). The functional significance of the hyporheiczone in streams and rivers. Annu. Rev. Ecol. Syst. 29, 59.

Boulton, A.J., Scarsbrook, M.R., Quinn, J.M., and Burrell, G.P.(2011). Land-use effects on the hyporheic ecology of five smallstreams near Hamilton, New Zealand. New Zeal. J. Mar. Fresh.31, 609.

Bradley, P.M. (2008). Potential for biodegradation of con-taminants of emerging concern in streams systems. Proceedingsof the 2008 South Carolina Water Resources Conference. Charles-ton, SC: U. S. Geological Survey South Carolina Water ScienceCenter.

Bradley, P.M., Barber, L.B., Chapelle, F.H., Gray, J.L., Kolpin,Dana, W.K., and McMahon, P.B. (2009). Biodegradation of17b-estradiol, estrone and testosterone in stream sediments.Environ. Sci. Technol. 43, 1902.

Brar, S.K. (2011). Emerging contaminants—Small concentrations,big problem. Hydrol. Curr. Res. 2, 101.

Briggs, M.A., Gooseff, M.N., Arp, C.D., and Baker, M.A. (2009).A method for estimating surface transient storage parameterswith concurrent hyporheic storage. Water Resour. Res. 45,W00D27.

Briggs, M.A., Lautz, L.K., McKenzie, J.M., Gordon, R.P., andHare, D.K. (2012). Using high-resolution distributed temper-ature sensing to quantify spatial and temporal variability invertical hyporheic flux. Water Resour. Res. 48, W02527.

Brown, L.R., Cuffney, T.F., Coles, J.F., Fitzpatrick, F., McMahon,G., Steuer, J., Bell, A.H., and May, J.T. (2009). Urban streamsacross the USA: Lessons learned from studies in 9 metropol-itan areas. J. N. Am. Benthol. Sci. 28, 1051.

Cardenas, M.B. (2008). Surface–ground water interface geo-morphology leads to scaling of residence times. Geophys. Res.Lett. 35, L08402.

Cardenas, M.B., and Wilson, J.L. (2007). Dunes turbulent eddiesand interfacial exchange with permeable sediments. WaterResour. Res. 43, W08412.

Cardenas, M.B., Wilson, J.L., and Zlotnik, V.A. (2004). Impact ofheterogeneity bed forms and stream curvature on subchannelhyporheic exchange. Water Resour. Res. 40, W08307.

Cardinale, B.J. (2011). Biodiversity improves water qualitythrough niche partitioning. Nature 472, 86.

Carey, R.O., and Migliaccio, K.W. (2009). Contribution ofwastewater treatment plant effluents to nutrient dynamics inaquatic systems: A review. Environ. Manage. 44, 205.

Chapelle, F.H., Bradley, P.M., McMahon, P.B., Kaiser, K., andBenner, R. (2012). Dissolved oxygen as an indicator of bio-available dissolved organic carbon in groundwater. GroundWater 50, 230.

Charbonneau, R., and Resh, V.H. (1992). Strawberry Creek onthe University of California, Berkeley campus: A case historyof urban stream restoration. Aquat. Conserv. 2, 293.

494 LAWRENCE ET AL.

Chen, B., Nam, S., Westerhoff, P.K., Krasner, S.W., and Amy, G.(2009). Fate of effluent organic matter and DBP precursors inan effluent-dominated river: A case study of wastewater im-pact on downstream water quality. Water Res. 43, 1755.

Chen, X. (2011). Fluctuation of hyporheic zone thickness due toinflow and outflow across the water-sediment interface. WaterResour. Environ. Prot. (ISWREP) 1, 388.

Choi, J., Harvey, J.W., and Conklin, M.H. (2000). Characterizingmultiple timescales of stream and storage zone interaction thataffect solute fate and transport in streams. Water Resour. Res.36, 1511.

Chu, C., Jones, N.E., Mandrak, N.E., Piggott, A.R., and Minns,C.K. (2008). The influence of air temperature, groundwaterdischarge, and climate change on thermal diversity of streamfishes in southern Ontario watersheds. Can. J. Fish. Aquat. Sci.65, 297.

Cirmo, C.P., and McDonnell, J.J. (1997). Linking the hydrologicand biogeochemical controls of nitrogen transport in near-stream zones of temperate-forested catchments: A review.J. Hydrol. 1, 88.

Claessens, L., and Tague, C. (2009). Transport-based method forestimating in-stream nitrogen uptake at ambient concentrationfrom nutrient addition experiments. Limnol. Oceanogr. Methods7, 811.

Claessens, L., Tague, C.L., Groffman, P.M., and Melack, J.M.(2010a). Longitudinal assessment of the effect of concentrationon stream N uptake rates in an urbanizing watershed. Bio-geochemistry 98, 63.

Claessens, L., Tague, C.L., Groffman, P.M., and Melack, J.M.(2010b). Longitudinal and seasonal variation of stream N uptakein an urbanizing watershed: Effect of organic matter, streamsize, transient storage and debris dams. Biogeochemistry 98, 45.

Clinton, S.M., Edwards, R.T., and Findlay, S.E.G. (2010).Exoenzyme activities as indicators of dissolved organic mattercomposition in the hyporheic zone of a floodplain river.Freshw. Biol. 55, 1603.

Coleman, J.C., Miller, M.C., and Mink, F.L. (2011). Hydrologicdisturbance reduces biological integrity in urban streams.Environ. Monit. Assess. 172, 663.

Collins, B.M., Sobczak, W.V., and Colburn, E.A. (2007). Sub-surface flowpaths in a forested headwater stream harbor adiverse macroinvertebrate community. Wetlands 27, 319.

Conant, B., Jr., Cherry, J.A., and Gillham, R.W. (2004). A PCEgroundwater plume discharging to a river: Influence of thestreambed and near-river zone on contaminant distributions.J. Contam. Hydrol. 73, 249.

Conn, K.E., Lowe, K.S., Drewes, J.E., Hoppe-Jones, C., and Tu-cholke, M.B. (2010). Occurrence of pharmaceuticals and com-sumer product chemicals in raw wastewater and septic tankeffluent from single-family homes. Environ. Eng. Sci. 27, 347.

Cousins, I.T., Kong, D., and Vestergren, R. (2011). Reconcilingmeasurements and modeling studies of the sources and fate ofperfluorinated carboxylates. Environ. Chem. 8, 339.

Crenshaw, C.L., Grimm, N.B., Zeglin, L.H., Sheibley, R.W., Dahm,C.N., and Ershall, A.D.P. (2010). Dissolved inorganic nitrogendynamics in the hyporheic zone of reference and human-alteredsouthwestern U.S. streams. Fund. Appl. Limnol. 176, 391.

Daigger, G.T. (2007). Wastewater management in the 21st cen-tury. ASCE J. Environ. Eng. 133, 671.

Datry, T. (2012). Benthic and hyporheic invertebrate assem-blages along a flow intermittence gradient: Effects of durationof dry events. Freshw. Biol. 57, 563.

David, A., Perrin, J.L., Rosain, D., Rodier, C., Picot, B., andTournoud, M.G. (2011). Implications of two in-stream pro-

cesses in the fate of nutrients discharged by sewage systeminto a temporary river. Environ. Monit. Assess. 181, 491.

Del Rosario, R.B., and Resh, V.H. (2000). Invertebrates in inter-mittent and perennial streams: Is the hyporheic zone a refugefrom drying? J. N. Am. Benthol. Sci. 19, 680.

Dennehy, K.F., Litke, D.W., Tate, C.M., Qi, S.L., McMahon, P.B.,Bruce, B.W., Kimbrough, R.A., and Heint, J.S. (1998). WaterQuality in the South Platte River Basin, Colorado, Nebraska,and Wyoming, 1992–95. U.S. Geological Survey Circular 1167.Denver, CO: USGS.

Ding, W.H., Wu, J., Semadeni, M., and Reinhard, M. (1999).Occurrence and behavior of wastewater indicators in theSanta Ana River and the underlying aquifers. Chemosphere 39,1781.

Dole-Olivier, M.-J. (2011). The hyporheic refuge hypothesis re-considered: A review of hydrological aspects. Mar. Freshw.Res. 62, 1281.

Dosskey, M.G., Vidon, P., Gurwick, N.P., Allan, C.J., Duval, T.P.,and Lowrance, R. (2010). The role of riparian vegetation inprotecting and improving chemical water quality in streams. J.Am. Water Resour. Assoc. 46, 261.

Durand, V., Aller, M.-F., Greswell, R.B., Whelan, J., Rivett, M.O.,MacKay, R., and Tellam, J.H. (2007). Natural attenuation po-tential of the urban hyporheic zone: Field experiment set-up,baseline monitoring and modeling for extraction test design.Presented at 2nd Switch Scientific Meeting, Tel Aviv, Israel,November 25–29.

Dzombak, D.A., and Morel, F.M.M., (1990). Surface ComplexationModelling- Hydrous Ferric Oxide. New York: John Wiley.

Ebersole, J.L., Liss, W.J., and Frissell, C.A. (1997). FORUM:Restoration of stream habitats in the western United States—Restoration as reexpression of habitat capacity. Environ.Manage. 21, 1.

Fellman, J.B., Hood, E., Edwards, R.T., and Jones, J.B. (2009).Uptake of allochthonous dissolved organic matter from soiland salmon in coastal temperate rainforest streams. Ecosystems12, 747.

Fellows, C.S., Valett, H.M., and Dahm, C.N. (2001). Whole-stream metabolism in two mountain streams: Contribution ofthe hyporheic zone. Limnol. Oceanogr. 46, 523.

Feral, D., Camann, M.A., and Welsh, H.H., Jr. (2005). Di-camptodon tenebrosus larvae within hyporheic zones of inter-mittent streams in California. Herpetol. Rev. 36, 26.

Fernald, A.G., Wigington, P.J., Jr., and Landers, D.H. (2001).Transient storage and hyporheic exchange along the Will-amette River, Oregon: Field measurements and model esti-mates. Water Resour. Res. 37, 1681.

Findlay, S.E. (1995). Importance of surface-subsurface exchangein stream ecosystems: the hyporheic zone. Limnol. Oceanogr.40, 159.

Findlay, S.E., Sinsabaugh, R.L., Sobczak, W.V., and Hoostal, M.,(2003). Metabolic and structural response to variation in sup-ply of dissolved organic matter. Limnol. Oceanogr. 48, 1608.

Findlay, S.E., Strayer, D., Goumbala, C., and Gould, K. (1993).Metabolism of streamwater dissolved organic carbon in shal-low hyporheic zone. Limnol. Oceanogr. 37, 1493.

Fischer, H., Kloep, F., Wilzcek, S., and Pusch, M.T. (2005). Ariver’s liver—Microbial processes within the hyporheic zoneof a large lowland river. Biogeochemistry 76, 349.

Fraser, B.G., and Williams, D.D. (1997). Accuracy and precisionin sampling hyporheic fauna. Can. J. Fish. Aquat. Sci. 54, 1135.

Gadd, G.M., and White, C.W. (1993). Microbial treatment ofmetal pollution—A working biotechnology? Trends Biotechnol.11, 353.

ACTIVE MANAGEMENT OF HYPORHEIC ZONES IN URBAN STREAMS 495

Gandy, C.J., Smith, J.W.N., and Jarvis, A.P. (2007). Attenuationof mining-derived pollutants in the hyporheic zone: A review.Sci. Total Environ. 373, 435.

Ge, Y., and Boufadel, M.C. (2006). Solute transport in multiple-reach experiments: Evaluation of parameters and reliability ofprediction. J. Hydrol. 323, 106.

Geza, M., Murray, K.E., and McCray, J.E. (2010a). Watershed scaleimpacts of nitrogen from onsite wastewater systems: Parametersensitivity and model calibration. J. Environ. Eng. 136, 926.

Geza, M., Murray, K.E., and McCray, J.E. (2010b). Model eval-uation of potential impacts of onsite wastewater systems onphosphorus in Turkey Creek Watershed. J. Environ. Qual. 39,1636.

Gibert, J., and Deharveng, L. (2002). Subterranean ecosystems: Atruncated functional biodiversity. Bioscience 52, 473.

Glassmeyer, S.T., Furlong, E.T., Kolpin, D.W., Cahill, J.D.,Zaugg, S.D., Werner, S.L., Meyer, M.T., and Kryak, D.D.(2005). Transport of chemical and microbial compoundsfrom known wastewater discharges: Potential for use as in-dicators of human fecal contamination. Environ. Sci. Technol.39, 5157.

Gobel, P., Dierkes, C., and Coldewey, W.G. (2007). Storm waterrunoff concentration matrix for urban areas. J. Contam. Hydrol.91, 26.

Gooseff, M.N. (2010). Defining hyporheic zones—advancing ourconceptual and operational definitions of where stream waterand groundwater meet. Geogr. Compass 4, 945.

Grant, S.B., Litton-Mueller, R.M., and Ahn, J.H. (2011). Mea-suring and modeling the flux of fecal bacteria across thesediment–water interface in a turbulent stream. Water Resour.Res. 47, W05517.

Grant, S.B., and Marusic, I. (2011). Crossing turbulent bound-aries: Interfacial flux in environmental flows. Environ. Sci.Technol. 45, 7107.

Grebel, J., Boehm, A., Mohanty, S., Torkelson, A., Sedlak, D.,Nelson, K., and Maxwell, R. (2013). Engineered infiltrationsystems for distributed urban stormwater reclamation. En-viron. Eng. Sci. 30, 437.

Grimaldi, C., and Chaplot, V. (1999). Nitrate depletion duringwithin-stream transport: Effects of exchange processes be-tween streamwater, the hyporheic and riparian zones. WaterAir Soil Pollut. 124, 95.

Grimm, N.B., Arrowsmith, J.R., Eisinger, C., Heffernan, J., Ma-cLeod, A., Lewis, D.B., Prashad, L., Rychener, T., Roach, J.W.,and Sheibley, R.W. (2004). Effects of urbanization on nutrientbiogeochemistry of aridland streams. Geophys. Monogr. Series153, 129.

Grimm, N.B., Sheibley, R.W., Crenshaw, C.L., Dahm, C.N.,Roach, W.J., and Zeglin, L.H. (2005). N retention and trans-formation in urban streams. J. N. Am. Benthol. Sci. 24, 626.

Grischek, T., Foley, A., Schoenheinz, D., and Gutt, B. (2002).Effects of interaction between surface water and groundwateron groundwater flow and quality beneath urban areas. InK.H.F. Howard, and R.G. Israfilov, Eds., Current Problems ofHydrogeology in Urban Areas, Urban Agglomeraates and In-dustrial Centres. Dordrecht, The Netherlands: Kluwer Aca-demic Publishers, pp. 201–219.

Groffman, P.M., Dorsey, A.M., and Mayer, P.M. (2005). N pro-cessing within geomorphic structures in urban streams. J. N.Am. Benthol. Soc. 24, 613.

Groning, J., Held, C., Garten, C., Claußnitzer, U., Kaschabek,S.R., and Schlomann, M. (2007). Transformation of diclofenacby the indigeneous microflora of river sediments and identi-fication of a major intermediate. Chemosphere 69, 509.

Gross, B., Montgomery-Brown, J., Naumann, A., and Reinhard,M. (2004). Occurrence and fate of pharmaceuticals and alkyl-phenol ethoxylate metabolites in an effluent-dominated riverand wetland. Environ. Toxicol. Chem. 23, 2074.

Haggard, B.E., Stanley, E.H., and Storm, D.E. (2005). Nutrientretention in a point-source–enriched stream. J. N. Am. Benthol.Soc. 24, 29.

Hancock, P.J., Boulton, A.J., and Humphreys, W.F. (2005).Aquifers and hyporheic zones: Towards an ecological under-standing of groundwater. Hydrogeol. J. 13, 98.

Harvey, J.W., and Fuller, C.C. (1998). Effect of enhanced man-ganese oxidation in the hyporheic zone on basin-scale geo-chemical mass balance. Water Resour. Res. 34, 623.

Harvey, J.W., Wagner, B.J., and Bencala, K.E. (1996). Evaluatingthe reliability of the stream tracer approach to characterizestream–subsurface water exchange. Water Resour. Res. 32,2441.

Hatch, C.E., Fisher, A.T., Revenaugh, J.S., Constantz, J., andRuehl, C. (2006). Quantifying surface water-groundwater in-teractions using time series of streambed thermal records:Method development. Water Resour. Res. 42, W10410.

Hawkins, C.P., Olson, J.R., and Hill, R.A. (2010). The referencecondition: Predicting benchmarks for ecological and water-quality assessments. J. N. Am. Benthol. Soc. 29, 312.

Heffernan, J.B., Sponseller, R.A., and Fisher, S.G. (2008). Con-sequences of a biogeomorphic regime shift for the hyporheiczone of a Sonoran Desert stream. Freshw. Biol. 53, 1954.

Hendricks, S.P. (1993). Microbial ecology of the hyporheic zone:A perspective intergrating hydrology and biology. J. N. Am.Benthol. Soc. 12, 70.

Hester, E.T., and Gooseff, M.N. (2010). Moving beyond thebanks: Hyporheic restoration is fundamental to restoringecological services and functions of streams. Environ. Sci. Tech.44, 1521.

Higgins, C.P., Sharp, J.O., Sepulvado, J.G., Littrell, B.J., O’Con-nor, G., Snyder, E., and McAvory, D. (2010). Trace organicchemicals in biosolids-amended soils—State-of-the-science-review. Water Environ. Res. Found. SRSK5T09.

Hill, A.R., Labadia, C.F., and Sanmugadas, K. (1998). Hyporheiczone hydrology and nitrogen dynamics in relation to thestreambed topography of a N-rich stream. Biogeochemistry 42,285.

Hockin, S.L., and Gadd, G.M. (2003). Linked redox precipitationof sulfur and selenium under anaerobic conditions by sulfate-reducing bacterial biofilms. Appl. Environ. Microbiol. 69, 7063.

Holmes, R.M., Jones, J.B., Jr., Fisher, S.G., and Grimm, N.B.(1996). Denitrification in a nitrogen-limited stream ecosystem.Biogeochemistry 33, 125.

Hoehn, E., Plumlee, M.H., and Reinhard, M. (2007). Naturalattenuation of potential downwelling streams for per-fluorochemicals and other emerging contaminants. Water Sci.Technol. 56, 59.

Hsu, B.-M., Huang, C.-C., Chen, J.-S., Chen, N.-H., and Huang,J.-T. (2011). Comparison of potentially pathogenic free-livingamoeba hosts by Legionella spp. in substrate-associated bio-films and floating biofilms from spring environments. WaterRes. 45, 5171.

Imai, A., Fukushima, T., Matsushige, K., Kim, Y.H., and Choi, K.(2002). Characterization of dissolved organic matter in efflu-ents from wastewater treatment plants. Water Res. 36, 859.

Ingendahl, D., Borchardt, D., Saenger, N., and Reichert, P.(2009). Vertical hydraulic exchange and the contribution ofthyporheic community respiration to whole ecosystem respi-ration in the River Lahn (Germany). Aquat. Sci. 71, 399.

496 LAWRENCE ET AL.

Jarvie, H.P., Mortimer, R.J.G., Palmer-Felgate, E.J., Quinton, K.S.,Harman, S.A., and Carbo., P. (2008). Measurement of solublereactive phosphorus concentration profiles and fluxes in river-bed sediments using DET gel probes. J. Hydrol. 350, 261.

Jasper, J., Nguyen, M., Jones, J., Ismail, N.S., Sedlak, D.L., Nelson,K.N., Sharp, J., and Luthy, R. (2013). Unit process wetlands fortreatment of municipal wastewater effluent: A review. Environ.Eng. Sci. 30, 421.

Kasahara, T., and Wondzell, S.M. (2003). Geomorphic controlson hyporheic exchange flow in mountain streams. Water Re-sour. Res. 39, 1005.

Kasahara, T., Datry, T., Mutz, M., and Boulton, A.J. (2009).Treating causes not symptoms: Restoration of surface–groundwater interactions in rivers. Mar. Freshw. Res. 60, 976.

Kasahara, T., and Hill, A.R. (2006). Effects of riffle-step restora-tion on hyporheic zone chemistry in N-rich lowland streams.Can. J. Fish. Aquat. Sci. 63, 120.

Kasahara, T., and Hill, A.R. (2008). Modeling the effects oflowland stream restoration projects on stream–subsurfacewater exchange. Ecol. Eng. 32, 310.

Kaushal, S.S., Likens, G.E., Jaworski, N.A., Pace, M.L., Sides,A.M., Seekell, D., Belt, K.T., Secor, D.H., and Wingate, R.L.(2010). Rising stream and river temperatures in the UnitedStates. Front. Ecol. Environ. 8, 461.

Kemp, P., Sear, D., Collins, A., Naden, P., and Jones, I. (2011).The impacts of fine sediment on riverine fish. Hydrol. Process.25, 1800.

Kennedy/Jenks Consultants. (2011). Willamette River TMDL Eastand West TMDL Alternatives Due Diligence Evaluation. Preparedfor City of Corvallis, Oregon, 15 November 2011. Available at:www.ci.corvallis.or.us/downloads/pw/Corvallis_EastWest_DueDiligence_Report_111511_complete_w-figures%20&%20appendices.pdf (accessed April 30, 2012).

Kinouchi, T., Yagi, H., and Miyamoto, M. (2007). Increase instream temperature related to anthropogenic heat input fromurban wastewater. J. Hydrol. 335, 78.

Kondolf, G.M., and Micheli, E.R. (1995). Evaluating stream res-toration projects. Environ. Manage. 19, 1.

Krause, S., Hannah, D.M., and Fleckenstein, J.H. (2009). Hy-porheic hydrology: Interactions at the groundwater-surfacewater interface. Hydrol. Process. 23, 2103.

Krause, S., Hannah, D.M., Fleckenstein, J.H., Heppell, C.M.,Kaeser, D., Pickup, R., Pinay, G., Robertson, A.L., and Wood,P.J. (2011). Inter-disciplinary perspectives on processes in thehyporheic zone. Ecohydrology 4, 481.

Kujawa-Roeleveld, K., Schuman, E., Grotenhuis, T., Kragic, D.,Mels, A., and Zeeman, G. (2008). Biodegradability of humanpharmaceutically active compounds (PhAC) in biologicalsystems treating source separated wastewater streams. ThirdSWITCH Scientific Meeting, Belo Horizone, Brazil.

Kunkel, U., and Radke, M. (2008). Biodegradation of acidicpharmaceuticals in bed sediments: Insight from a laboratoryexperiment. Environ. Sci. Technol. 42, 7273.

Kunkel, U., and Radke, M., (2011). Reactive tracer test to eval-uate the fate of pharmaceuticals in rivers. Environ. Sci. Technol.45, 6296.

Landmeyer, J.E., Bradley, P.M., Trego, D.A., Hale, K.G., andHaas II, J.E. (2010). MTBE, TBA, and TAME attenuation indiverse hyporheic zones. Ground Water 48, 30.

Lapworth, D.J., Gooddy, D.C., Allen, D., and Old, G.H. (2009).Understanding groundwater, surface water and hyporheiczone biogeochemical processes in a chalk catchment usingfluorescence properties of dissolved and colloidal organicmatter. J. Geophys. Res. 114, G00F02.

Lapworth, D.J., Gooddy, D.C., and Jarvie, H.P. (2011). Un-derstanding phosphorus mobility and bioavailability in thehyporheic zone of a chalk stream. Water Air Soil Pollut. 218,213.

Larsen, S., and Ormerod, S.J. (2010). Low-level effects of inertsediments on temperate stream invertebrates. Freshw. Biol. 55,476.

Lautz, L.K., and Fanelli, R.M. (2008). Seasonal biogeochemicalhotspots in the streambed around restoration structures. Bio-geochemistry, 91, 85.

Lautz, L.K., and Siegel, D.I. (2006). Modeling surface andground water mixing in the hyporheic zone using MODFLOWand MT3D. Adv. Water Resour. 29, 1618.

Lauver, L., and Baker, L.A. (2000). Mass balance for wastewaternitrogen in the central Arizona–Phoenix ecosystem. Water Res.34, 2754.

Lawrence, J.E., Lunde, K.B., Mazor, R.D., Beche, L.A., McElravy,E.P., and Resh, V.H. (2010). Long-term macroinvertebrate re-sponses to climate change: implications for biological assess-ment in Mediterranean-climate streams. J. N. Am. Benthol. Soc.29, 1424.

Lear, G., and Lewis, G.D. (2009). Impact of catchment land useon bacterial communities within stream biofilms. Ecol. Indic. 9,848.

Lewandowski, J., Putschew, A., Schwesig, D., Neumann, C., andRadke, M. (2011). Fate of organic micropollutants in the hy-porheic zone of a eutrophic lowland stream: Results of apreliminary field study. Sci. Total Environ. 10, 1824.

Li, J.L., Johnson, S.L., and Sobota, J.B. (2011). Three responses tosmall changes in stream temperature by autumn-emergingaquatic insects. J. N. Am. Benthol. Sci. 30, 474.

Liess, M., and von der Ohe, P.C. (2005). Analyzing effects ofpesticides on invertebrate communities in streams. Environ.Toxicol. Chem. 24, 954.

Lijklema, L. (1980). Interaction of orthophosphate with iron(III)and aluminum hydroxides. Environ. Sci. Technol. 14, 537.

Lim, M.H., Snyder, S.A., and Sedlak, D.L. (2008). Use of biode-gradable dissolved organic carbon (BDOC) to assess the po-tential for transformation of wastewater-derived contaminantsin surface waters. Water Res. 42, 2943.

Lin, A.Y.C., Plumlee, M.H., and Reinhard, M. (2006). Naturalattenuation of pharmaceuticals and alkylphenol poly-ethocylate metabolites during river transport: Photochemicaland biological transformation. Environ. Toxicol. Chem. 25, 1458.

Loffler, D., Rombke, J., Meller, M., and Ternes, T. (2005). En-vironmental fate of pharmaceuticals in water/sediment sys-tems. Environ. Sci. Technol. 39, 5209.

Lopez-Rodrıguez, M.J., Tierno de Figueroa, J.M., Fenoglio, S.,Bo, T., and Alba-Tercedor, J. (2009). Life strategies of 3 Per-lodidae species (Plecoptera) in a Mediterranean seasonalstream in southern Europe. J. N. Am. Benthol. Sci. 28, 611.

Louhi, P., Ovaska, M., Maki-Petays, A., Erkinaro, J., andMuotka, T. (2011). Does fine sediment constrain salmonidalevin development and survival? Can. J. Fish. Aquat. Sci. 68,1819.

Maier, H.S., and Howard, K.W.F. (2011). Influence of oscillatingflow on hyporheic zone development. Ground Water 49, 830.

Malard, F., and Hervant, F. (1999). Oxygen supply and the ad-aptations of animas in groundwater. Freshw. Biol. 41, 1.

Malard, F., Tockner, K., Dole-Olivier, M.-J., and Ward, J.V.(2002). A landscape perspective of surface–subsurface hydro-logical exchanges in river corridors. Freshw. Biol. 47, 621.

Malcolm, I.A., Middlemas, C.A., Soulsby, C., Middlemas, S.J.,and Youngson, A.F. (2010). Hyporheic zone processes in a

ACTIVE MANAGEMENT OF HYPORHEIC ZONES IN URBAN STREAMS 497

canalised agricultural stream: Implications for salmonid em-bryo survival. Fund. Appl. Limnol. 176, 319.

Martin, J.D. (2009). Sources and preparation of data for assessingtrends in concentrations of pesticides in streams of the UnitedStates, 1992–2006. U.S. Geological Survey Scientific In-vestigations Report 2009-5062, p. 41.

Mayer, P.M., Groffman, P.M., Striz, E.A., and Kaushal, S.S.(2010). Nitrogen dynamics at the groundwater–surface waterinterface of a degraded urban stream. J. Environ. Qual. 39, 810.

McBride, G.B., and Tanner, C.C. (1999). Modelling biofilm ni-trogen transformations in constructed wetland mesocosmswith fluctuating water levels. Ecol. Eng. 14, 93.

McDermott, M.J., Robertson, A.L., Shaw, P.J., and Milner, A.M.(2010). The hyporheic assemblage of a recently formed streamfollowing deglaciation in Glacier Bay, Alaska, USA. Can. J.Fish. Aquat. Sci. 67, 304.

McMahon, P.B., Tindall, J.A., Collins, J.A., and Lull, K.J. (1995).Hydrologic and geochemical effects on oxygen uptake inbottom sediments of an effluent-dominated river. Water Res.31, 2561.

Meakins, N.C., Bubb, J.M., and Lester, J.N. (1994). The fate andbehaviour of organic micropollutants during wastewatertreatment processes: A review. Int. J. Environ. Pollut. 4, 27.

Mendez, G.O., and Belitz, K. (2002). Groundwater/Surface WaterInteractions AWRA Summer Specialty Conference. Available at:http://infohouse.p2ric.org/ref/33/32611.pdf (accessed April30, 2012).

Mermillod-Blondin, F., Mauclaire, L., and Montuelle, B. (2005).Use of slow filtration columns to assess oxygen respiration,consumption of dissolved organic carbon, nitrogen transfor-mations, and microbial parameters in hyporheic sediments.Water Res. 39, 1687.

Meyer, E.I., Niepagenkemper, O., Molls, F., and Spanhoff, B.(2008). An experimental assessment of the effectiveness ofgravel cleaning operations in improving hyporheic waterquality in potential salmonid spawning areas. River Res. Appl.24, 119.

Mikkelson, K.M., Maxwell, R.M., Sharp, J.O., Stednick, J., andMcCray, J.E. (2013). Mountain pine beetle infestation impacts:Modeling water and energy budgets at the hillslope scale.Ecohydrol. J. 6, 64.

Monteiro, S.C., and Boxall, A.B.A. (2010). Occurrence and fate ofhuman pharmaceuticals in the environment. Rev. Environ.Contam. T. 202, 53.

Mueller, M., Pander, J., and Geist, J. (2011). The effects of weirson structural stream habitat and biological communities. J.Appl. Ecol. 48, 1450.

Mulholland, P.J., Marzolf, E.R., Webster, J.R., Hart, D.R., andHendricks, S.P. (1997). Evidence that hyporheic zones increaseheterotrophic metabolism and phosphorus uptake in foreststreams. Limnol. Oceanogr. 42, 443.

Mulholland, P.J., Tank, J.L., Sanzone, D.M., Wollheim, W.M.,Peterson, B.J., Webster, J.R., and Meyer, J.L. (2000). Nitrogencycling in a forest stream determined by a 15N tracer addition.Ecol. Monogr. 70, 471.

Murphy, S.F., Verplanck, P.L., and Barber, L.B., Eds. (2000).Comprehensive water quality of the Boulder CreekWatershed, under high-flow and low-flow conditions, 2000.U.S. Geological Survey Water Resources Investigation Report03-4045. Washington, DC.

Musson, S.E., Campo, P., Tolaymat, T., Suidan, M., and Town-send, T.G. (2010). Assessment of the anaerobic degradation ofsix active pharmaceutical ingredients. Sci. Total Environ. 408,2068.

Naegeli, M.W., and Uehlinger, U. (1997). Contribution of thehyporheic zone to ecosystem metabolism in a prealpinegravel-bed river. J. N. Am. Benthol. Sci. 16, 794.

Nelson, K.C., and Palmer, M.A. (2007). Stream temperaturesurges under urbanization and climate change: data, models,and responses. J. Am. Water Resour. Assoc. 43, 440.

Nogaro, G., Datry, T., Mermillod-Blondin, F., Descloux, S., andMontuelle, B. (2010). Influence of streambed sediment clog-ging on microbial processes in the hyporheic zone. Freshw.Biol. 55, 1288.

O’Connor, B.L., Hondzo, M., and Harvey, J.W. (2010). Predictivemodeling of transient storage and nutrient uptake: Implica-tions for stream restoration. J. Hydraul. Eng. 136, 1018.

Olden, J.D., and Naiman, R.J. (2010). Incorporating thermal re-gimes into environmental flows assessments: Modifying damoperations to restore freshwater ecosystem integrity. Freshw.Biol. 55, 86.

Olden, J.D., Poff, N.L., and Bestgen, K.R. (2006). Life-historystrategies predict fish invasions and extirpations in the Col-orado River Basin. Ecol. Monogr. 76, 25.

Olsen, D.A., Matthaei, C.D., and Townsend, C.R. (2010). Effectsof a depositional flood event on the hyporheos of a NewZealand stream. Fund. Appl. Limnol. 176, 337.

Oram, J.J., McKee, L.J., Werme, C.E., Connor, M.S., Oros, D.R.,Grace, R., and Rodigari, F. (2008). A mass budget of poly-brominated diphenyl ethers in San Francisco Bay, CA. En-viron. Int. 34, 1137.

Paillex, A., Doledec, S., Castella, E., and Merigoux, S. (2009).Large river floodplain restoration: Predicting species richnessand trait responses to the restoration of hydrological connec-tivity. J. Appl. Ecol. 46, 250.

Palmer, M.A., Lettenmaier, D.P., Poff, N.L., Postel, S.L., Richter,B., and Warner, R. (2009). Climate change and river ecosys-tems: Protection and adaptation options. Environ. Manage. 44,1053.

Palmer-Felgate, E.J., Mortimer, R.J.G., Krom, M.D., and Jarvie,H.P. (2010). Impact of point-source pollution on phosphorusand nitrogen cycling in stream-bed sediments. Environ. Sci.Technol. 44, 908.

Paria, S. (2008). Surfactant-enhanced remediation of organiccontaminated soil and water. Adv. Colloid Interface 138, 24.

Paul, M.J., and Meyer, J.L. (2001). Streams in the urban land-scape. Ann. Rev. Ecol. Syst. 32, 333.

Parc Cientific de Barcelona (PCB). (2001). Stream Reach Manage-ment: An Expert System. Available at: www.pcb.ub.edu/streames/sites/erpe.htm#location (accessed April 30, 2012).

Petrunic, B.M., MacQuarrie, K.T.B., and Al, T.A. (2005). Re-ductive dissolution of Mn oxides in river-recharged aquifers:A laboratory column study. J. Hydrol. 301, 163.

Pinay, G., O’Keefe, T.C., Edwards, R.T., and Naiman, R.J. (2009).Nitrate removal in the hyporheic zone of a salmon river inAlaska. River Res. Appl. 25, 367.

Plumlee, M.H., Larabee, J., and Reinhard, M. (2008). Per-fluorochemicals in water reuse. Chemosphere 72, 1541.

Plumlee, M.H., and Reinhard, M. (2007). Photochemical atten-uation of N-nitrosodimethylamine (NDMA) and other ni-trosoamines in surface water. Environ. Sci. Technol. 41, 6170.

Poff, N.L., Allan, J.D., Bain, M.B., Karr, J.R., Prestegaard, K.L.,Richter, B.D., Sparks, R.E., and Stromberg, J.C. (1997). Thenatural flow regime. Bioscience 47, 769.

Poff, N.L., Richter, B.D., Arthington, A.H., Bunn, S.E., Naiman,R.J., Kendy, E., Acreman, M., Apse, C., Bledsoe, B.P., Freeman,M.C., Henriksen, J., Jacobson, R.B., Kennen, J.G., Merritt,D.M., O’Keeffe, J.H., Olden, J.D., Rogers, K., Tharme, R.E., and

498 LAWRENCE ET AL.

Warner, A. (2010). The ecological limits of hydrologic alter-ation (ELOHA): A new framework for developing regionalenvironmental flow standards. Freshw. Biol. 55, 147.

Pohlon, E., Marxsen, J., and Kusel, K. (2010). Pioneering bacterialand algal communities and potential extracellular enzymeactivities of stream biofilms. FEMS Microbiol. Ecol. 71, 364.

Pollen-Bankhead, N., and Simon, A. (2010). Hydrologic andhydraulic effects of riparian root networks on streambankstability: Is mechanical root-reinforcement the whole story?Geomorphology 116L, 353.

Power, M.E., Sun, A., Parker, G., Dietich, W.E., and Wooton, J.T.(1995). Hydraulic food-chain models. Bioscience 45, 159.

Radke, M., Ulrich, H., Wurm, C., and Kunkel, U. (2010). Dy-namics and attenuation of acidic pharmaceuticals along a riverstretch. Environ. Sci. Technol. 44, 2968.

Rau, G.C., Anderson, M.S., McCallum, A.M., and Acworth, R.I.(2010). Analytical methods that use natural heat as a tracer toquantify surface water–groundwater exchange, evaluated us-ing field temperature records. Hydrogeol. J. 18, 1093.

Rayne, S., and Forest, K. (2009). Pewrfluoroalkyl sulfonic andcarboxylic acids: A critical review of physicochemical prop-erties, levels and patterns in waters and wastewaters, andtreatment methods. J. Environ. Sci. Health Part A 44, 1145.

Resh, V.H., Brown, A., Covich, A.P., Gurtz, M.E., Li, H.W.,Minshall, W., Reice, S., Sheldon, A.L., Wallace, J.B., andWissmar, R. (1988). The role of disturbance theory in streamecology. J. N. Am. Benthol. Soc. 7, 433.

Resh, V.H., and Rosenberg, D.M. (2010). Recent trends in life-history research on benthic macroinvertebrates. J. N. Am.Benthol. Sci. 29, 207.

Reynolds, S.K., Jr., and Benke, A.C. (2012). Chironomid pro-duction along a hyporheic gradient in contrasting streamtypes. Freshw. Sci. 31, 167.

Richter, B.D., Baumgartner, J.V., Wigington, R., and Braun, D.P.(1997). How much water does a river need? Freshw. Biol. 37,231.

Robertson, A.L., and Wood, P.J. (2010). Ecology of the hyporheiczone: origins, current knowledge and future directions. Fund.Appl. Limnol. 176, 279.

Rosenberry, D.O., and Pitlick, J. (2009). Local-scale spatial andtemporal variability of seepage in a shallow gravel-bed river.Hydrol. Process. 23, 3306.

Rosenberry, D.O., Klos, P.Z., and Neal, A. (2012). In situ quan-tification of spatial and temporal variability of hyporheic ex-change in static and mobile gravel-bed rivers. Hydrol. Process.26, 604.

Roy, J.W., and Bickerton, G. (2011). Toxic groundwater con-taminants: An overlooked contributor to urban stream syn-drome? Environ. Sci. Technol. 46, 729.

Runkel, R.L. (1998). One-dimensional transport with inflow andstorage (OTIS): A solute transport model for streams andrivers. U.S. Geologic Survey Water Resources Investigations Re-port 98-4018. U.S. Geological Survey, Denver, Colorado.

Ryan, R.J., and Boufadel, M.C. (2006a). Incomplete mixing in asmall urban stream. J. Environ. Manage. 81, 50.

Ryan, R.J., and Boufadel, M.C. (2006b). Evaluation of streambedhydraulic conductivity heterogeneity in an urban watershed.Stoch. Environ. Res. Risk Assess. 21, 309.

Ryan, R.J., and Boufadel, M.C. (2007). Lateral and longitudinalvariation of hyporheic exchange along a mountain stream.Environ. Sci. Technol. 41, 4221.

Ryan, R.J., Larson, P.C., and Welty, C. (2011). The logistics andmechanics of conducting tracer injection experiments in urbanstreams. Urban Ecosyst. 14, 87.

Ryan, R.J., Welty, C., and Larson, P. (2010). Variation in surfacewater–groundwater exchange with land use in an urbanstream. J. Hydrol. 392, 1.

Sabater, S., Gausch, H., Romanı, A., and Munoz, I. (2002). Theeffect of biological factors on the efficiency of river biofilms inimproving water quality. Hydrobiologia 469, 149.

Sabater, S., and Tockner, K. (2010). Effects of hydrologic alter-ations on the ecological quality of river ecosystems. Handb.Environ. Chem. 8, 15.

Salehin, M., Packman, A.I., and Paradis, M. (2004). Hyporheicexchange with heterogeneous streambeds: Laboratory exper-iments and modeling. Water Resour. Res. 40, W11504.

Sarriquet, P.E., Bordenave, P., and Marmonier, P. (2007). Effectsof bottom sediment restoration on interstitial habitat charac-teristics and benthic macroinvertebrate assemblages in aheadwater stream. River Res. Appl. 23, 815

Sawyer, A.H., Cardenas, M.B., Bomar, A., and Mackey, M.(2009). Impact of dam operations on hyporheic exchange inthe riparian zone of a regulated river. Hydrol. Process. 23, 2129.

Schindler, J.E., and Krabbenhoft, D.P. (1998). The hyporheiczone as a source of dissolved organic carbon and carbon gasesto a temperate forested stream. Biogeochemistry 43, 157.

Schipper, L.A., Cooper, A.B., Harfoot, C.G., and Dyck, W.J.(1993). Regulators of denitrification in an organic riparian soil.Soil Biol. Biochem. 25, 925.

Schmidt, C., Bayer-Raich, M., and Schirmer, M. (2006). Char-acterization of spatial heterogeneity of groundwater-streamwater interactions using multiple depth streambed tempera-ture measurements at the reach scale. Hydrol. Earth Syst. Sci.10, 849.

Schultz, M.M., Furlong, E.T., Kolpin, D.W., Werner, S.L.,Schoenfuss, H.L., Barber, L.B., Blazer, V.K., Norris, D.O., andVajda, A.M. (2010). Antidepressant pharmaceuticals in twoU.S. effluent-impacted streams: occurrence and fate in waterand sediment, and selective uptake in fish neural tissue. En-viron. Sci. Technol. 44, 1918.

Scordo, E.B., and Moore, R.D. (2009). Transient storage processesin a steep headwater stream. Hydrol. Process. 23, 2671.

Sedlak, D.L., Gray, J.L., and Pinkston, K.E. (2000). Under-standing micro contaminants in recycled water. Environ. Sci.Technol. 34, 509.

Smedley, P.L., and Kinniburgh, D.G. (2002). A review of thesource, behavior, and distribution of arsenic in natural water.Appl. Geochem. 17, 517.

Smith, E.L. (2000). Pre-Research Survey of Municipal NPDES Dis-chargers in the Arid and Semi-Arid West. Prepared for PimaCounty Wastewater Management Department by URS Corp.,San Francisco, CA.

Smith, S.D., Devitt, D.A., Sala, A., Cleverly, J.R., and Busch, D.E.(1998). Water relations of riparian plants from water desertregions. Wetlands 18, 687.

Song, J., Chen, X., and Cheng, C. (2010). Observation of bio-turbation and hyporheic flux in streambeds. Front. Environ.Sci. Eng. China 4, 340.

Soulsby, C., Malcolm, I.A., Tetzlaff, D., and Youngson, A.F.(2009). Seasonal and inter-annual variability in hyporheicwater quality revealed by continuous monitoring in a salmonspawning stream. River Res. Appl. 25, 1304.

Sponseller, R.A., Grimm, N.B., Boulton, A.J., and Sabo, J.L.(2010). Responses of macroinvertebrate communities to long-term flow variability in a Sonoran Desert stream. GlobalChange Biol. 16, 2891.

Stelzer, R.S., Bartsch, L.A., Richardson, W.B., and Strauss, E.A.(2011). The dark side of the hyporheic zone: Depth profiles of

ACTIVE MANAGEMENT OF HYPORHEIC ZONES IN URBAN STREAMS 499

nitrogen and its processing in stream sediments. Freshw. Biol.56, 2021.

Sternecker, K., and Geist, J. (2010). The effects of stream sub-stratum composition on the emergence of salmonid fry. Ecol.Freshw. Fish 19, 537.

Stranko, S.A., Hilderbrand, R.H., and Palmer, M.A. (2012).Comparing the fish and benthic macroinvertebrate diversity ofrestored urban streams to reference streams. Restor. Ecol. 20,747.

Strauss, E.A., Mitchell, N.L., and Lamberti, G.A. (2002). Factorsregulating nitrification in aquatic sediments: effects of organiccarbon, nitrogen availability, and pH. Can. J. Fish. Aquat. Sci.59, 554.

Stubbington, R., Wood, P.J., and Reid, I. (2011). Spatial vari-ability in the hyporheic zone refugium of temporary streams.Aquat. Sci. 73, 499.

Stumm, W., and Morgan, J.J. (1996). Aquatic Chemistry—ChemicalEquilibria and Rates in Natural Waters, 3rd edition. New York:Wiley-Interscience.

Sullivan, S.M.P., and Watzin, M.C. (2010). Towards a functionalunderstanding of the effects of sediment aggradation onstream fish condition. River Res. Appl. 26, 1298.

Ternes, T.A., Joss, A., and Siegrist, H. (2008). Scrutinizingpharmaceuticals and personal care products in wastewatertreatment. Environ. Sci. Technol. 38, 392.

Tharme, R.E. (2003). A global perspective on environmental flowassessment: emerging trends in the development and appli-cation of environmental flow methodologies for rivers. RiverRes. Appl. 19, 397.

Thurman, E.M. (1985). Organic Geochemistry of Natural Waters.Dordrecht, Germany: Kluwer Academic.

Toet, S., Huibers, L.H.F.A., van Logtestijn, R.S.P., and Verhoe-ven, J.T.A. (2003). Denitrification in the periphyton associatedwith plant shoots and in the sediment of a wetland systemsupplied with sewage treatment plant effluent. Hydrobiologia501, 29.

Tomlinson, M., and Boulton, A.J. (2010). Ecology and manage-ment of subsurface groundwater dependent ecosystems inAustralia—A review. Mar. Freshw. Res. 61, 936.

Tonina, D., and Buffington, J.M. (2009). A three-dimensionalmodel for analyzing the effects of salmon redds on hyporheicexchange and egg pocket habitat. Can. J. Fish. Aquat. Sci. 66,2157.

Toran, L.T., Hughes, B., Nyquist, J.E., and Ryan, R.J. (2012). Usinghydrogeophysics to monitor change in hyporheic flow aroundstream restoration structures. Environ. Eng. Geosci. 18, 83.

Triska, F.J., Duff, J.H., and Avanzino, R.J. (1993). The role ofwater exchange between a stream channel and its hyporheiczone in nitrogen cycling at the terrestrial-aquatic interface.Hydrobiologia 251, 167.

Treese, S., Meixner, T., and Hogan, J.F. (2009). Clogging of aneffluent dominated semiarid river: a conceptual model ofstream aquifer interactions. J. Am. Water Resour. Assoc. 45,1047.

Tufenkji, N., Ryan, J.N., and Elimelech, M. (2002). Bank filtra-tion: a simple technology may inexpensively clean up poor-quality raw surface water. Environ. Sci. Technol. 36, 422.

Uzarski, D.G., Stricker, C.A., Burton, T.M., King, D.K., andSteinman, A.D. (2004). The importance of hyporheic sedimentrespiration in several mid-order Michigan rivers: Comparisonbetween methods in estimates of lotic metabolism. Hydro-biologia 518, 47.

Vajda, A.M., Barber, L.B., Gray, J.L., Lopez, E.M., Woodling,J.D., and Norris, D.O. (2008). Reproductive disruption in fish

downstream from an estrogenic wastewater effluent. Environ.Sci. Technol. 42, 3407.

van den Akker, B., Holmes, M., Pearce, P., Cromar, N.J., andFallowfield, H.J. (2011). Structure of nitrifying biofilms in ahigh-rate trickling filter designed for potable water pre-treatment. Water Res. 45, 3489.

Vaux, W.G. (1968). Intragravel flow and interchange of water ina streambed. Fish. Bull. 66, 479.

Verplanck, P.L., Taylor, H.E., Nordstrom, D.K., and Barber, A.B.(2005). Aqueous stability of gadolinium in surface waters re-ceiving sewage treatment plant effluent, Boulder Creek, Col-orado. Environ. Sci. Technol. 39, 6923.

Wagner, F.H., and Beisser, C. (2005). Does carbon enrichmentaffect hyporheic invertebrates in a gravel stream? Hydro-biologia 544, 189.

Walsh, C.J., Roy, A.H., Feminella, J.W., Cottingham, P.D.,Groffman, P.M., and Morgan II, R.P. (2005). The urban streamsyndrome: Current knowledge and the search for a cure. J. N.Am. Benthol. Soc. 24, 706.

Walters, A.W. (2011). Resistance of aquatic insects to a low-flowdisturbance: Exploring a trait-based approach. J. N. Am. Ben-thol. Sci. 30, 346.

Ward, A.S., Gooseff, M.N., and Johnson, P.A. (2011). How cansubsurface modifications to hydraulic conductivity be de-signed as stream restoration structures? Analysis of Vaux’sconceptual models to enhance hyporheic exchange. WaterResourc. Res. 47, W08512.

Ward, J.V. (1989). The four-dimensional nature of lotic ecosys-tems. J. N. Am. Benthol. Sci. 8, 2.

Webb, B.W., Hannah, D.M., Moore, R.D., Brown, L.E., andNobilis, F. (2008). Recent advances in stream and river tem-perature research. Hydrol. Process. 22, 902.

Wesener, M.D., Merritt, R.W., Campbell, E.Y., Benbow, M.E.,Norberg, E.C., and Edwards, R.T. (2011). Hyporheic andbenthic macroinvertebrate communities in glacial, clearwater,and brownwater streams in Alaska. Pan-Pac. Entomol. 87, 145.

Westerhoff, P., and Anning, D. (2000). Concentrations andcharacteristics of inorganic carbon in surface water in Arizona:Influence of urbanization. J. Hydrol. 236, 202.

White, D.S. (1993). Perspectives on defining and delineatinghyporheic zones. J. N. Am. Benthol. Sci. 12, 61.

Williams, D.D. (1989). Towards a biological and chemical defi-nition of the hyporheic zone in two Canadian rivers. Freshw.Biol. 22, 189.

Williams, D.D., Febria, C.M., and Wong, J.C.Y. (2010). Ecotonaland other properties of the hyporheic zone. Fund. Appl. Limnol.176, 349.

Withers, P.J.A., and Jarvie, H.P. (2008). Delivery and cycling ofphosphorus in rivers: A review. Sci. Total Environ. 400, 379.

Wondzell, S.M., LaNier, J., and Haggerty, R. (2009). Evaluationof alternative groundwater flow models for simulating hy-porheic exchange in a small mountain stream. J. Hydrol. 364,142.

Wood, P.J., Boulton, A.J., and Stubbington, R. (2010). Is thehyporheic zone a refugium for aquatic macroinvertebratesduring severe low flow conditions? Fund. Appl. Limnol. 176,377.

Worrall, F., Harriman, R., Evans, C.D., Watts, C.D., Adamson, J.,Neal, C., Tipping, E., Burt, T., Grieve, I., Manteith, D., Naden,P.S., Nisbet, T., Reynolds, B., and Stevens, P. (2004). Trends indissolved organic carbon in UK rivers and lakes. Biogeochem-istry 70, 369.

Writer, J.H., Barbert, L.B., Ryan, J.N., and Bradley, P.M. (2011).Biodegradation and attenuation of steroidal hormones and

500 LAWRENCE ET AL.

alkylphenols by stream biofilms and sediments. Environ. Sci.Technol. 45, 4370.

Yamada, H., and Nakamura, F. (2009). Effects of fine sediment ac-cumulation on the redd environment and the survival rate of masusalmon (Oncorhynchus masou) embryos. Landsc. Ecol. Eng. 5, 169.

Young, B.A., Norris, R.H., and Sheldon, F. (2011). Is the hy-porheic zone a refuge for macroinvertebrates in drying pe-rennial streams? Mar. Freshw. Res. 62, 1373.

Zarnetske, J.P., Haggerty, R., Wondzell, S.M., and Baker, M.A.(2011). Dynamics of nitrate production and removal as afunction of residence time in the hyporheic zone. J. Geophys.Res. 116, G01025.

Zushi, Y., Takeda, T., and Masunaga, S. (2008). Existence ofnonpoint source of perfluorinated compounds and theirloads in the Tsurumi River basin, Japan. Chemosphere 71,1566.

ACTIVE MANAGEMENT OF HYPORHEIC ZONES IN URBAN STREAMS 501