Top Banner
Atmospheric Environment 39 (2005) 2309–2322 Atmospheric concentrations and deposition of organochlorine pesticides in the US Mid-Atlantic region Rosalinda Gioia a, , John H. Offenberg a , Cari L. Gigliotti b , Lisa A. Totten a , Songyan Du a , Steven J. Eisenreich c a Department of Environmental Sciences, Rutgers, The State University of New Jersey, 14 College Farm Road, New Brunswick, NJ 08901, USA b Department of Chemistry, Brookdale College, 765 Newman Spring Road, Lincroft, NJ 07738, USA c Joint Research Centre: Inland and Marine Waters, TP 290, Ispra, Varese I-21020, Italy Received 1 September 2004; accepted 7 December 2004 Abstract Organochlorine pesticides (OCPs) were measured in the atmosphere over the period January 2000–May 2001 at six locations as part of New Jersey Atmospheric Deposition Network (NJADN). Gas phase, particle phase and precipitation concentrations of 22 OCP species, including chlordanes, DDTs, HCHs, endosulfan I and II, aldrin and diedrin, were measured. OCPs are found predominantly in the gas phase in all seasons, representing over 95% of the total air concentrations. Most of the pesticides measured display highest concentrations at urban sites (Camden and New Brunswick), although in many cases the differences in geometric mean concentrations are not statistically significant. The relationship of gas-phase partial pressure with temperature was examined using the Clausius–Clapeyron equation; significant temperature dependencies were found for all OCPs, except aldrin. Atmospheric depositional fluxes (gas absorption into water+dry particle deposition+wet deposition) to the New York–New Jersey Harbor Estuary of selected OCPs were estimated at NJADN sites. Atmospheric concentrations of dieldrin, aldrin and the HCHs are similar to those measured by the Integrated Atmospheric Deposition Network (IADN) in the Great Lake Region. In contrast, concentrations of DDTs, chlordanes and heptachlor are higher in the Mid-Atlantic compared to the Great Lakes, suggesting that the New York–New Jersey Harbor Estuary receives higher fluxes of these chemicals than the Great Lakes. r 2005 Elsevier Ltd. All rights reserved. Keywords: Organochlorine pesticides; Atmospheric deposition; Volatilization; Clausius–Clapeyron equation; Enthalpy of the environment 1. Introduction Organochlorine pesticides (OCPs) were widely used in North America until the 1970s when many were banned. Information on the global distribution of these com- pounds has increased during the last 20 years, due to their persistence in the environment, which allows them to be transported over great distances. As a result, OCPs ARTICLE IN PRESS www.elsevier.com/locate/atmosenv 1352-2310/$ - see front matter r 2005 Elsevier Ltd. All rights reserved. doi:10.1016/j.atmosenv.2004.12.028 Corresponding author. Current address: Department of Environmental Science, Institute of Environmental and Natural Science, Lancaster University, Lancaster LA1 4YQ, UK. Tel.: +44 1524 593974; fax: +44 1524 593985. E-mail address: [email protected] (R. Gioia).
14

Atmospheric concentrations and deposition of organochlorine pesticides in the US Mid-Atlantic region

Apr 21, 2023

Download

Documents

Welcome message from author
This document is posted to help you gain knowledge. Please leave a comment to let me know what you think about it! Share it to your friends and learn new things together.
Transcript
Page 1: Atmospheric concentrations and deposition of organochlorine pesticides in the US Mid-Atlantic region

ARTICLE IN PRESS

1352-2310/$ - se

doi:10.1016/j.at

�CorrespondEnvironmental

Science, Lanca

Tel.: +441524

E-mail addr

Atmospheric Environment 39 (2005) 2309–2322

www.elsevier.com/locate/atmosenv

Atmospheric concentrations and deposition of organochlorinepesticides in the US Mid-Atlantic region

Rosalinda Gioiaa,�, John H. Offenberga, Cari L. Gigliottib, Lisa A. Tottena,Songyan Dua, Steven J. Eisenreichc

aDepartment of Environmental Sciences, Rutgers, The State University of New Jersey, 14 College Farm Road,

New Brunswick, NJ 08901, USAbDepartment of Chemistry, Brookdale College, 765 Newman Spring Road, Lincroft, NJ 07738, USA

cJoint Research Centre: Inland and Marine Waters, TP 290, Ispra, Varese I-21020, Italy

Received 1 September 2004; accepted 7 December 2004

Abstract

Organochlorine pesticides (OCPs) were measured in the atmosphere over the period January 2000–May 2001 at six

locations as part of New Jersey Atmospheric Deposition Network (NJADN). Gas phase, particle phase and

precipitation concentrations of 22 OCP species, including chlordanes, DDTs, HCHs, endosulfan I and II, aldrin and

diedrin, were measured. OCPs are found predominantly in the gas phase in all seasons, representing over 95% of the

total air concentrations. Most of the pesticides measured display highest concentrations at urban sites (Camden and

New Brunswick), although in many cases the differences in geometric mean concentrations are not statistically

significant. The relationship of gas-phase partial pressure with temperature was examined using the Clausius–Clapeyron

equation; significant temperature dependencies were found for all OCPs, except aldrin. Atmospheric depositional fluxes

(gas absorption into water+dry particle deposition+wet deposition) to the New York–New Jersey Harbor Estuary of

selected OCPs were estimated at NJADN sites. Atmospheric concentrations of dieldrin, aldrin and the HCHs are

similar to those measured by the Integrated Atmospheric Deposition Network (IADN) in the Great Lake Region. In

contrast, concentrations of DDTs, chlordanes and heptachlor are higher in the Mid-Atlantic compared to the Great

Lakes, suggesting that the New York–New Jersey Harbor Estuary receives higher fluxes of these chemicals than the

Great Lakes.

r 2005 Elsevier Ltd. All rights reserved.

Keywords: Organochlorine pesticides; Atmospheric deposition; Volatilization; Clausius–Clapeyron equation; Enthalpy of the

environment

e front matter r 2005 Elsevier Ltd. All rights reserve

mosenv.2004.12.028

ing author. Current address: Department of

Science, Institute of Environmental and Natural

ster University, Lancaster LA1 4YQ, UK.

593974; fax: +44 1524 593985.

ess: [email protected] (R. Gioia).

1. Introduction

Organochlorine pesticides (OCPs) were widely used in

North America until the 1970s when many were banned.

Information on the global distribution of these com-

pounds has increased during the last 20 years, due to

their persistence in the environment, which allows them

to be transported over great distances. As a result, OCPs

d.

Page 2: Atmospheric concentrations and deposition of organochlorine pesticides in the US Mid-Atlantic region

ARTICLE IN PRESSR. Gioia et al. / Atmospheric Environment 39 (2005) 2309–23222310

have also been found in the Artic (Bidleman et al., 2002;

Hung et al., 2002) and have been shown to bioaccumu-

late and biomagnify in animals of higher trophic levels

such as whales, seals, polar bears and humans (Halsal et

al., 1998; Kucklick and Baker, 1998; Muir et al., 1996).

Although the use of these compounds has been

eliminated in most western industrialized countries,

many of them are still in use in Africa and South and

Central America (Alegria et al., 2000; Janunten et al.,

2000). Several long-term studies have shown that the

concentration of OCPs in the atmosphere of the

Northern Hemisphere has diminished over the last 10

years (Cortes et al., 1998, 1999). However, 20 years after

they were banned, OCPs remain ubiquitous in the

environment. Therefore, transport from countries south

of the USA border and re-emission from soils due to

Fig. 1. Sites of the NJADN. OCP concentrations are reported for Jer

Delaware Bay.

past usage support measurable levels of these com-

pounds in the Northern Hemisphere (Hoff et al.,

1992a, b; Janunten et al., 2000).

The New Jersey Department of Environmental

Protection (NJDEP) conducted an historical review of

OCP use in New Jersey in order to provide insight into

the type and the possible geographical extent of

potential residual pesticide contamination (NJDEP,

1999). DDT (including DDD and DDE), aldrin and

dieldrin were considered ‘‘pesticides of concern’’ because

these groups are among the most widely used pesticides

over the last century. Heptachlor (HEPT), chlordanes

and endosulfan (I and II) were also included in this

study (NJDEP, 1999).

Atmospheric concentrations of OCPs have been

measured in Alabama and South Carolina, where the

sey City, New Brunswick, Sandy Hook, Pinelands, Camden and

Page 3: Atmospheric concentrations and deposition of organochlorine pesticides in the US Mid-Atlantic region

ARTICLE IN PRESSR. Gioia et al. / Atmospheric Environment 39 (2005) 2309–2322 2311

past usage of these compounds was widespread, and in

the Great Lakes Region, where the atmosphere can

serve as a pathway for the delivery of these pollutants to

water and terrestrial surfaces (Bidleman et al., 1998a, b;

Cortes et al.,1998; Harner et al., 1999; Hillery et al.,

1998; Janunten et al., 2000; Leone et al., 2001;

McConnell et al., 1996).

However, little research has been conducted on

chlorinated pesticides in the air of the Mid-Atlantic

region of the United States. A research and monitoring

network, NJADN (New Jersey Atmospheric Deposition

Network), was established in 1997 in order to assess the

fate and transport of persistent organic pollutants,

including PAHs (polycyclic aromatic hydrocarbons),

PCBs (polychlorinated biphenyls) and OCPs, in coastal,

urban, suburban and rural locations in the Mid-Atlantic

area (Brunciak et al., 2001; Gigliotti et al., 2000, 2002;

Van Ry et al., 2002; Offenberg et al., 2004; Totten et al.,

2004).

This work will use data from NJADN (i) to assess the

occurrence and spatial variability of OCP concentra-

tions, (ii) to investigate the seasonality associated with

the OCP concentrations, (iii) to estimate atmospheric

deposition fluxes (wet+dry+gas absorption) of selected

OCPs to the New York–New Jersey Harbor Estuary

(Fig. 1).

2. Methodology

2.1. Sampling and site characterization

Air samples were collected at six different locations in

New Jersey representing different land-use regimes:

suburban (New Brunswick), coastal/marine (Sandy

Hook and Delaware Bay), urban/industrial (Jersey City

and Camden) and rural/forested (Pinelands). Detailed

descriptions of these sites can be found in Brunciak et al.

(2001), Gigliotti et al. (2000, 2002), Van Ry et al. (2002),

Offenberg et al. (2004) and Totten et al. (2004).

Twenty-four hour integrated air samples were col-

lected every 12 days over the period January 2000–May

2001. At each site, modified Tisch Environmental High

volume air samplers operated at a calibrated air flow

rate of �500Lmin�1 in order to collect both particulate

and gas phase of SOCs including PCBs, PAHs and

OCPs.

2.2. Sample processing

Details of the analysis can be found in Gioia (2004)

and the publications above. The particulate phase was

captured on pre-combusted quartz fiber filters (QFF)

and the gas phase was captured on polyurethane foam

(PUF). QFFs were pre- and post-weighed under

temperature and humidity control to determine total

suspended particulates (TSP).

All samples were spiked with a surrogate standard

containing d10-anthracene, d10-fluoranthene and d12-

benzo[e]pyrene for PAHs and PCB14, PCB23, PCB65

and PCB166 for PCBs. Each was extracted in Soxhlet

apparatus for 24 h, the PUFs in petroleum ether and

QFFs with dichloromethane. The extracts were concen-

trated using rotoevaporation and nitrogen evaporation.

Wet-only precipitation collectors were employed

(MIC, Ontario) to collect precipitation in 0.212m2

stainless-steel funnels that drained through a column

containing XAD-2 resin. The volume of filtered rain was

measured at the end of each 24-day sampling period and

varied from 2.7 to 20L. XAD rain samples were

extracted in acetone:hexane (1:1 by volume) in Soxhlet

apparatus for 24 h after the addition of surrogates to

gauge analytical recoveries. The extracts were liquid–li-

quid extracted with 60mL deionized water. The aqueous

fractions were back-extracted with 3� 50mL hexane in

separatory funnels containing 1 g sodium chloride. The

samples were then concentrated by rotoevaporation and

treated in the same manner as air and water samples.

Each sample was eluted on a 3% water-deactivated

alumina column. Fraction 1, which contains PCBs and

some chlorinated pesticides (o,p0-DDE, HEPT, aldrin,

p,p0-DDE, o,p0-DDT), was eluted with 13mL hexane.

Fraction 2, which contains PAHs and some chlorinated

pesticides (trans- and cis-chlordanes (TC and CC), trans-

and cis-nonachlors (TN and CN), dieldrin, p,p0-DDT,

o,p0-DDT), was eluted with 15mL 2:1 dichlorometha-

ne:hexane. Fractions were reduced in volume to

�0.5mL under N2 gas. Both fractions were spiked with

internal standard consisting of PCB30 and PCB204,

while Fraction 2 was also spiked with internal standard

consisting of d10-phenanthrene, d10-pyrene, and d12-

benzo[a]pyrene. Both fractions were analyzed for OCPs.

PCB204 was used as the internal standard for OCPs

analysis. Surrogate recoveries were calculated for Frac-

tion 1 by using PCB166 and ranged from 90% to 105%;

however, it was not possible to calculate surrogate

recoveries for the F2 fractions due to the poor response

of PAHs in NCI mode. However, surrogate recoveries

for PAHs ranged from 70% to 98% when analyzed in

the gas chromatography/mass spectrometry/electron

impact (GC/MS/EI) mode. Due to the high surrogate

recoveries, analyte amount were not corrected for

surrogate extraction efficiencies.

The OCPs were analyzed using gas chromatography/

mass spectrometry in negative chemical ionization (GC/

MS-NCI) operating in selective ion monitoring mode

using a Hewlett Packard 6890 GC coupled to a Hewlett

Packard 5973 Mass Selective Detector and a

30m� 0.25mm i.d., J&W Scientific 122-5062 DB-5

(5% diphenyl-dimethylpolysiloxane) capillary column

with a film thickness of 0.25mm. The ratio of the

Page 4: Atmospheric concentrations and deposition of organochlorine pesticides in the US Mid-Atlantic region

ARTICLE IN PRESSR. Gioia et al. / Atmospheric Environment 39 (2005) 2309–23222312

primary ion to the secondary ion in samples was taken

as 720% of the ratio that was in the calibration

standard. The mass of each OCP was determined and

quantified based on relative response factor, which was

calculated from five standards of varying concentration

(1 –100 ngml�1).

2.3. Quality assurance/quality control

Field blanks, laboratory blanks and matrix spikes

were included at a rate of 15%, 10% and 5%,

respectively. No OCPs were detectable in the laboratory

and field blanks at the MDLs. Therefore, instrumental

detection limits were calculated for each analyte using a

series of seven dilutions of the lowest (1 ngml�1)

calibration standard; the series of dilutions were

analyzed seven times and the MDL was taken as t

value� standard deviation of the data set of the

solution where the response of the instrument is no

longer linear. MDLs ranged from 1.2 pg for o,p0-DDT

to 78 pg for p,p0-DDD. At a typical sample volume of

600m3, instrumental detection limits would be 0.002–

0.13 pgm�3 for o,p0-DDT and p,p0-DDD, respectively.

Recovery for target analyte (HEPT, p,p0-DDT, p,p0-

DDE and o,p0-DDT) spike samples (matrix spike) range

from about 70% for p,p0-DDT and 112% for HEPT.

Replicates of samples were also analyzed at the end of

each run in order to calculate precision. The precision

was calculated as the square root of the sum of variance

of all replicates divided by the mean concentrations of

pesticides and was 10–15% for all compounds. Split

PUFs were used to quantify breakthrough of gas-phase

OC pesticides. The bottom half accounted for 1575%(n ¼ 4) of the total mass with the greatest breakthrough

occurring for a-HCH (40–50%). Breakthrough was

lower for other pesticides: 3–25% for g-HCH and

1–2% for chlordanes. Since a-HCH breakthrough was

found to be relatively high, a-HCH concentrations

reported in this study are minimum estimates.

2.4. Framework for deposition calculations

High volume air samplers employed in this study

collected a bulk dry atmospheric particle sample. The

dry deposition flux is calculated by applying a single

deposition velocity (Vd) to the bulk particle phase

Fdry ¼ VdCpart, (1)

where Fdry is the flux (ngm�2 day�1) and Cpart is the

average particle concentration of the chemical (ngm�3).

We selected a value for the Vd of 0.5 cm s�1 that

represents the disproportionate influence that large

particles have on atmospheric deposition, especially in

urbanized and industrialized regions (Franz et al., 1998;

Pirrone et al., 1995a, b) Some Great Lake studies

employed a deposition velocity of 0.2 cm s�1 (Hillery

et al., 1998; Miller et al., 2001). Thus, the choice of Vdrepresents the largest source of error in the calculation

of dry deposition flux. Moreover, Vd may differ between

sites due to changes in particle characteristics. The error

in the dry deposition fluxes is therefore conservatively

estimated to be about 760%.Wet deposition fluxes (Fwet) are estimated as follows:

Fwet ¼ CWVMP, (2)

where CWVM is the volume-weighted mean concentra-

tion of the chemical in precipitation and P is the

precipitation intensity (m yr�1). The standard error of

the volume-weighted mean concentration (see below)

dominates the error in Fwet.

The air–water exchange flux was estimated using the

modified two-layer resistance model developed by

Withman (1923), which assumes that the rate of transfer

is controlled by the compound’s ability to diffuse across

two thin stagnant films at the air–water interface, the

water film and the air film.

The overall flux calculation is defined by

F ¼ vawðCw � Ca=KawÞ, (3)

where F is the flux (ngm�2 day�1), vaw (mday�1) is the

overall mass transfer coefficient and (Cw�Ca/Kaw)

describes the concentration gradient (ngm�3). Cw(ngm�3) is the dissolved phase concentration of the

compound in water. Ca (ngm�3) is the gas-phase

concentration of the compound in air which is divided

by the dimensionless Henry’s Law Constant, Kaw. The

net flux is divided into volatilization (Fgas, vol) and

absorption (Fgas, abs) term as follows:

Fgas; vol ¼ vawCw, (4)

Fgas; abs ¼ vawCa=Kaw. (5)

Only the absorptive gas flux (Eq. (5)) is calculated in

this study, using the gas-phase OCP concentrations

measured at the land-based sites, because Cw was not

available. Fgas, abs was calculated separately for each day

of sample collection, and the results are averaged to

yield an annual average of Fgas, abs. For all pesticides, the

wind speed was assumed to be constant at 5m s�1

(Totten et al., 2004). Salinity is assumed to be zero, but

an increase in Kaw due to salinity will decrease the gas

absorption flux. The error in the Fgas, abs (Eq. (5))

depends on the uncertainty in Vaw, which is about

740% (Nelson et al., 1998; Bamford et al., 1999, 2001),

and the uncertainty in Ca, which is about 725%. Thepropagated error in Fgas, abs is thus 47%. This figure does

not encompass the error in Kaw because it is assumed to

be a systematic, not random, error (Nelson et al., 1998;

Bamford et al., 1999, 2001). Henry’s Law Constants

were taken from Mackay et al. (1999) and Iwata et al.

Page 5: Atmospheric concentrations and deposition of organochlorine pesticides in the US Mid-Atlantic region

ARTICLE IN PRESSR. Gioia et al. / Atmospheric Environment 39 (2005) 2309–2322 2313

(1995) for chlordanes, DDTs and aldrin/dieldrin, from

Suntio et al. (1988) for endosulfans, and from Sahsuvar

et al. (2003) for HCHs. Enthalpies of vaporization for all

compounds except HCH came from Hinckley et al.

(1990); enthalpies of vaporization for HCHs came from

Sahsuvar et al. (2003).

3. Results and discussion

Tables 1–3 summarize the OCP measurements. b-HCH, d-HCH and p,p0-DDD were never detected.

Table 1

Average (avg) and geometric mean (GM) concentrations of gas-phase

six NJADN sites

Camden (Jan

00–May 01)

(N ¼ 43)

Delaware Bay

(Mar 00–May 01)

(N ¼ 34)

Jersey City (Ja

00–Jan 01)

(N ¼ 30)

Avg GM Avg GM Avg GM

HEPT 304 269 32 21 228 200

HEPTX 45 26 11 9.3 8.1 3.

Dieldrin 75 51 15 10 8.2 2.

Aldrin 77 71 1.3 1.1 3.8 6.

Endosulfan I 102 22 32 11 13 3.

Endosulfan II 1.8 3.6 Not

detected

1.9 0.

a-HCH 96 73 37 34 23 16

g-HCH 158 80 34 20 22 12

S-Chlordanes 518 354 119 87 114 72

S-DDTs 133 92 39 19 31 25

Table 2

Average (avg) and geometric mean (GM) organochlorine pesticide

particulate (TSP) (mgm�3) and total number of the samples analyzed

Camden (Jan

00–May 01) (N ¼ 30)

Delaware Bay (Mar

00–May 01) (N ¼ 21)

Jersey

00–Ja

Avg GM Avg GM Avg

TSP 54 51 55 49 52

Oxychlordane 0.12 0.12

Dieldrin 2.8

Aldrin 0.23

Endosulfan I 2.0 1.1 0.84 0.49 0.98

Endosulfan II 4.2 1.5 2.3 1.0 1.6

Endosulfan

sulfate

1.7 0.76 1.0 0.68 0.81

S-Chlordanes 6.4 4.3 0.96 0.77 6.1

S-DDTs 3.4 3.6 1.7 1.5 4.9

Blank spaces indicate that the pesticide was never detected in the par

3.1. S-Chlordanes

Technical chlordane was mostly used as an agricul-

tural pesticide on corn and citrus, for home lawns and

gardens and as a termiticide in house foundations. The

most abundant components in technical mixture are TC,

CC, TN, CN, b-chlordane and HEPT (Henry et al.,

1994).

Gas-phase concentrations of S-chlordanes averaged83 to 518 pgm�3 for the different sites, while concentra-

tions of particle-phase Schlordanes averaged

1–10 pgm�3, such that o�5% of all atmospheric S-chlordane was typically present in the particle phase.

For both particle- and gas-phase chlordanes, highest

OCPs (pgm�3) and total number of the samples analyzed (N) at

n New Brunswick

(Jan 00–Jan 01)

(N ¼ 26)

Pinelands (Jan

00–May 01)

(N ¼ 43)

Sandy Hook (Jan

00–Jan 01)

(N ¼ 25)

Avg GM Avg GM Avg GM

115 101 48 41 65 41

5 19 12 14 16 1.8 1.5

3 36 30 38 22 9.7 3.4

9 28 23 1.1 2.5 1.2 4.6

1 168 21 59 14 15 4.1

7 10 2.6 10 8.4 2 0.76

84 62 55 49 19 16

65 33 48 26 18 8.1

474 223 127 83 83 45

237 106 31 16 53 22

concentrations (pgm�3) in the particle phase, total suspended

(N) at five NJADN sites

City (Jan

n 01) (N ¼ 29)

New Brunswick (Jan

00–Jan 01) (N ¼ 24)

Pinelands (Jan

00–May 01) (N ¼ 30)

GM Avg GM Avg GM

48 47 43 27 25

0.22 0.17 0.063 0.080

2.7

6.6

0.89 0.50 0.25 0.89 0.50

1.2 2.1 0.82 2.2 0.80

0.52 1.0 0.50 1.4 0.59

4.2 4.1 2.4 1.1 0.78

10 2.8 2.0 0.58 0.47

ticle phase at that site.

Page 6: Atmospheric concentrations and deposition of organochlorine pesticides in the US Mid-Atlantic region

ARTICLE IN PRESS

Table 3

Volume-weighted mean concentrations (VWM, pgL�1), standard error of the mean (SEM, pgL�1) for pesticides in precipitation

samples and total number of the samples analyzed at NJADN sites

Camden (Jan 00–May 01)

(N ¼ 12)

Jersey City (Jan 00–Jan 01)

(N ¼ 13)

Pinelands (Jan 00–May 01)

(N ¼ 30)

VWM SEM VWM SEM VWM SEM

Endosulfan I 154 57 46 21 153 82

Endosulfan II 788 327 293 130 283 80

Endosulfan sulfate 444 140 219 70 378 68

Oxychlordane 3.2 1.2 3.3 1.6 8.3 5.9

S-DDTsa 190 75 367 184 58 19

S-Chlordanes 173 40 182 55 66 8.3

a4,40-DDT was the only DDT compound detected in precipitation samples.

R. Gioia et al. / Atmospheric Environment 39 (2005) 2309–23222314

concentrations were typically observed at Camden and

New Brunswick, while Delaware Bay displays the lowest

S-chlordane concentrations. Particle-phase S-chlordaneconcentrations were also high at Jersey City. In the

precipitation phase, VWM concentrations were lowest

at Pinelands and significantly higher at Jersey City and

Camden.

High atmospheric S-chlordanes levels are likely to beassociated with past pesticide use. High S-chlordanelevels at New Brunswick are therefore not surprising, as

this site is located nearby fields used in agricultural

research. High OCP levels observed in Camden may be

due to its location downwind of rich agricultural areas in

eastern Pennsylvania, and to use and storage in the

urban area. Similarly, Harner et al. (2004) and Eitzer

et al. (2003) reported high air concentration of chlordane

in urban areas, possibly resulting from previous usage of

chlordane in house foundations.

Gas-phase S-chlordanes at most of the New Jersey

sites were similar to concentrations measured in Toronto

over the period July–October 2000 (about 90 pgm�3)

(Harner et al., 2004), in Alabama during 1996–1997

(98 pgm�3) (Janunten et al., 2000) and in South

Carolina during 1994–1995 (180 pgm�3) (Bidleman

et al., 1998a, b). In contrast, concentrations of S-chlordanes were generally lower (averagingo40 pgm�3)

at remote sites around the Great Lakes during

1996–1998 (Buehler et al., 2001), in the Cornbelt region

of the US during 1996–1997 (56 pgm�3) (Leone et al.,

2001) and around Lake Baikal in Russia during June

1991 (5 pgm�3) (McConnell et al., 1996). S-Chlordanesare higher in the Mid-Atlantic region than in the Great

Lakes area, suggesting that the New York–New Jersey

Harbor Estuary receives higher inputs of these chemicals

per unit area than the Great Lakes (Table 4).

Atmospheric deposition of chlordanes is dominated

by gas absorption, with dry and wet deposition

comprising o10% of the total inputs. Given the

estimated 47% uncertainty in the gas absorption fluxes,

significant differences in fluxes are apparent. Although

differences in gas absorption between the sites may not

be statistically significant, it is reasonable to assume that

higher atmospheric concentrations will lead to higher

deposition fluxes (although differences in particle size

distributions, wind speeds and salinity of receiving

waters can shift spatial deposition patterns).

3.2. HEPT and heptachlor epoxides (HEPTX)

HEPT is an organochlorine cyclodiene insecticide,

first isolated from technical chlordane in 1946. During

the 1960s and 1970s, it was used primarily in termite, ant

and soil insect control in seed grains and on crops, as

well as well as in the home. An important metabolite of

HEPT is HEPTX, which is an oxidation product formed

in many plant and animal species (Duerksen-Hughen

et al., 1993). HEPT is a moderately toxic compound in

EPA toxicity class II. Phase out of HEPT use began in

1978 and all but one use was cancelled in the US in 1988.

The only commercial use still permitted is for fire ant

control in power transformers. HEPT is still available

outside the US; HEPT is transformed in the environ-

ment by several pathways. Photolysis yields photohep-

tachlor and HEPTX, both of which are persistent,

bioaccumulative and toxic (Bidleman et al., 1998a, b).

Neither HEPT nor HEPTX were detected in the

particle phase. As with the chlordanes, HEPT and

HEPTX gas-phase concentrations were generally higher

at Camden and New Brunswick (Fig. 2a, Table 1),

although in contrast to the chlordanes, HEPT and

HEPTX levels were high at Jersey City as well. The

highest concentration at the urban sites Camden and

Jersey City may be associated with current use of HEPT.

These concentrations are well above the averages of 1.2

and 7.4 pgm�3 of HEPT and HEPTX, respectively,

measured by Integrated Atmospheric Deposition Net-

work (IADN) near Lake Superior in 1996 and 1997

(Buehler et al., 2001). Leone et al. (2001) measured an

Page 7: Atmospheric concentrations and deposition of organochlorine pesticides in the US Mid-Atlantic region

ARTICLE IN PRESS

Table 4

Mean (geometric mean) atmospheric concentrations (pgm�3) of organochlorine pesticides in the Southern US, in New Jersey and

Great Lakes Regions

S-HCHs S-Chlordanes Dieldrin S-DDTs

NJADN 2000

Camden 254 518 75 133

Delaware Bay 71 119 15 39

Jersey City 44 111 8.2 25

New Brunswick 149 474 36 237

Pinelands 103 127 38 31

Sandy Hook 36 79 9.3 42

IADN 1996–1998a

Chicago 130 130 130 71

Sturgeon Point 82 38 19 31

Sleeping Bear Dunes 99 14 15 11

Brule River 89 7.8 8.4 2.9

Eagle Harbor 96 8.6 8.8 4.4

Alabamab 142 98 38 10

South Carolinac 180

Lake Baikal Russiad 594 4.9 21

Combelt Region, USe 56 3.1 6.3

Indianaf 200

Arkansasf 160

Toronto (Gage)g 107 90 36 109

Egbert (rural/agricultural area near Toronto)g 102 51 76 305

a1996–1998 (Buehler et al., 2001).bJanuary–October 1996 and May 1997 (Janunten et al., 2000).c1994–1995 (Bidleman et al., 1998a, b).dJune 1991 (McConnell et al., 1996).eFall of 1996, spring/fall of 1997 (Leone et al., 2001).f2002–2003 (Hoh and Hites, 2004).gJuly–October 2000 (Harner et al., 2004).

R. Gioia et al. / Atmospheric Environment 39 (2005) 2309–2322 2315

average of 9.3 pgm�3 HEPT in the Cornbelt region of

the US during 1996 and 1997.

Gas absorption was significantly higher at Camden

and New Brunswick and lower at the Delaware Bay site.

Gas absorption of HEPTX was not calculated because a

reliable Kaw for this compound could not be found.

Because HEPT was below detection limit in all particle

and precipitation samples, dry and wet deposition

fluxes were not calculated and assumed to be negligible

(Table 5).

3.3. DDTs

DDT was first used to control disease-spreading

insects and then as a multipurpose insecticide. The peak

of production of DDT in the US was 82 million kg in

1962, and it was deregistered in 1972, except for public

health emergencies. (Faroon et al., 2002).

DDT compounds (p,p0-DDT p,p0-DDE p,p0-DDD

o,p0-DDT o,p0-DDE o,p0-DDD) were detected in most

of the air samples with the exception of p,p0-DDD,

which was consistently below the limits of detection. As

with the chlordanes, gas-phase S-DDT concentrationswere generally highest at New Brunswick and Camden

(Fig. 2a, Table 1). As for the chlordanes, high levels of

S-DDT in New Brunswick are not surprising because ofprobable past usage of OCP pesticides at this location. It

should, however, be noted that DDT was manufactured

at a plant on the Delaware River just south of

Philadelphia, which might explain the high DDT levels

measured in Camden. At the other four sites, the

geometric mean concentration of S-DDTs was about20 pgm�3. The levels of DDTs measured at the New

Jersey sites are similar to those reported for the Great

Lakes by IADN (Table 4) at Chicago and Sturgeon

Point, but higher than levels at the other IADN sites

(Buehler et al., 2001), with the exemption of Camden

and New Brunswick. In contrast, DDT concentrations

at these two sites are in the same range than those

measured in Toronto (Harner et al., 2004). Only p,p0-

DDT was detected in the particle and precipitation

phases (Tables 2 and 3), except at Jersey City, where all

Page 8: Atmospheric concentrations and deposition of organochlorine pesticides in the US Mid-Atlantic region

ARTICLE IN PRESS

Aldrin

0

100

200

300

400

500

Dieldrin

0

100

200

300

400

Endosulfan I + II

0

400

800

1200

Heptachlor

Gas

-pha

se c

once

ntra

tion

(pg

m-3

)

0

300

600

ΣChlordanes

0

500

1000

1500

2000

ΣDDTs

0

200

400

600

ΣHCHs

CC DB JC NB PL SH

0

(a)

300

600

900

CC DB JC NB PL

0

15

30

45

Part

icle

-pha

se c

once

ntra

tion

(in

pg/m

3 )

5

20

35

0

15

30

45

(b)

ΣChlordanes

Endosulfan I + II

ΣDDTs

(c) CC JC PL

0

2000

4000

6000

8000

10000

12000

Prec

ipita

tion

conc

entr

atio

n (p

g/L

)

1000

3000

5000

7000

500

1500

2500

3500Endosulfan I + II

ΣChlordanes

ΣDDTs

Fig. 2. (a) Box and whisker plot of (a) gas-phase, (b) particle-phase and (c) precipitation OCP concentrations; upper dot, upper error

bar, upper edge of box, lower edge of box, lower error bar and lower dot represent 95th, 90th, 75th, 25th, 10th and 5th percentile

concentrations, respectively. Within each box, mean and median concentrations are shown as dashed and solid lines, respectively.

R. Gioia et al. / Atmospheric Environment 39 (2005) 2309–23222316

of the DDT compounds were detected in the particle

phase. The fraction of the total atmospheric concentra-

tion of p,p0-DDT that was in the particle phase was

typically about 10%, although at Jersey City the fraction

averaged 58%.

The average DDT/DDE ratio is o1 at all sites,indicating that the relative lack of new sources of DDT

in the atmosphere at the five locations. The average

DDT/DDE ratio ranged from 0.4770.23 at Pinelands

to 0.1870.08 at New Brunswick. By comparison, ratiosof p,p0-DDT/p,p0-DDE are 0.2–0.8 at the Great Lakes

and 0.3–0.8 in the Canadian Artic (Alegria et al., 2000;

Cortes et al., 1998), and 0.17–1.45 in urban and rural

sites in Toronto areas (Harner et al., 2004).

Atmospheric deposition of DDTs is dominated by gas

absorption at most sites. As with chlordanes and HEPT,

high gas-phase concentration levels yield high gas

absorption. Wet+dry deposition constituted o10% of

Page 9: Atmospheric concentrations and deposition of organochlorine pesticides in the US Mid-Atlantic region

ARTICLE IN PRESS

Table 5

Atmospheric deposition fluxes (ngm�2 day�1) of select OCPs at NJADN sites

Gas absorption Dry particle deposition Wet deposition

CC DB JC NB PL SH CC DB JC NB PL CC JC PL

a-HCH 47 19 11 43 26 9.7

g-HCH 78 17 11 34 22 9.1

S-HCHs 125 36 23 77 48 19

HEPT 12 0.63 9 5.1 1.74 1.6

Dieldrin 31 6.3 3.5 15 15.0 4.2 1.2

Aldrin 10.8 0.0097 0.61 2.6 0.19 0.20 0.10

Endosulfan I 41 13 5.3 75 21 6.0 0.87 0.36 0.42 0.22 0.39 0.46 0.14 0.46

Endosulfan II 6.5 0 0.80 4.4 3.3 0.80 1.8 1.0 0.70 0.90 0.94 2.4 0.88 0.85

Endosulfan sulfate 0.72 0.44 0.35 0.43 0.60 1.3 0.66 1.1

Oxychlordane 0.052 0.097 0.027 0.0095 0.010 0.025

S-Chlordanes 172 38 38 147 39 27 2.8 0.42 2.7 1.8 0.46 0.52 0.55 0.20

S-DDTs 37 8.3 9.1 62 8.0 14 1.5 0.71 2.1 1.2 0.25 0.57 1.1 0.17

Blank spaces indicate that the pesticide was never detected in that phase at that site.

R. Gioia et al. / Atmospheric Environment 39 (2005) 2309–2322 2317

the total (gas+wet+dry) deposition (Table 5). At

Delaware Bay and Jersey City, dry particle and wet

deposition were significant, together comprising as much

as 34% of the total. S-DDTs fluxes are higher at Mid-Atlantic sites than at IADN sites, suggesting that the

New York–New Jersey Harbor Estuary receives higher

inputs of these chemicals per unit area than the Great

Lakes.

3.4. Aldrin/diedrin

Aldrin and diedrin were used extensively from the

1950s to the 1970s to fight insects on corn, cotton and

citrus crops, and also used as a termiticides. Aldrin is

rapidly metabolized to dieldrin in the environment. In

1972, however, the EPA approved aldrin and dieldrin

for killing termites. Use of aldrin and dieldrin continued

until 1987, when the manufacturer voluntarily cancelled

the registration of aldrin and dieldrin for killing termites

(Hanley et al., 2002).

Gas-phase aldrin concentrations were typically high-

est at Camden and New Brunswick (Fig. 2a, Table 1),

and lower at the other four sites. In contrast, gas-phase

dieldrin concentrations were more evenly distributed

throughout the state and were generally higher than

those of aldrin, consistent with the 1987 final ban of

aldrin use and the conversion of aldrin to dieldrin in the

environment. This suggests that aldrin is declining in the

atmosphere and being metabolized to dieldrin. Dieldrin

levels measured by NJADN were very similar to those

measured in the Cornbelt region of the US (Hoff et al.,

1992a), and by IADN at background sites around the

Great Lakes from 1996 to 1998 (Kallenborn, et al.,

1998) (Table 4). The lowest average concentrations

reported for both networks are about 8–9 pgm�3,

suggesting that this level may represent a regional

(northeastern US) background. Interestingly, this back-

ground concentration was observed at Jersey City, a

heavily urban and industrial area. Aldrin and dieldrin

were not detected in precipitation, and were detected in

the particle phase only at Jersey City (Table 3), where

the particle phase comprised 6% and 25% of the total

atmospheric concentrations of aldrin and dieldrin,

respectively.

Atmospheric deposition of both aldrin and dieldrin is

dominated by gas absorption (Table 5) at all sites, except

Jersey City, where dry deposition comprised about

20–30% of the total.

3.5. a and g-Hexachlorocyclohexanes

Technical HCH is a mixture of a-HCH (60–70%), b-HCH (2–12%) and g-HCH (10–15%) and several otherminor isomers (Iwata et al., 1993).

This technical mixture was used in the US until 1978.

Afterwards, it was substituted by the purified g-HCHisomer (lindane). Lindane has not been manufactured in

the US since 1977, but it is still imported into the

country. Its usage has been restricted by the US EPA to

certified individuals. Lindane is still used for seed

treatments (Walker et al., 1999).

Due to breakthrough of the HCHs during sampling,

the gas-phase concentrations presented here are mini-

mum estimates. Again, gas-phase concentrations of S-HCH (a-HCH+g-HCH) are generally highest at NewBrunswick and Camden (Fig. 2a, Table 1), although

differences between the sites are not statistically

significant. The HCHs were not detected in the particle

or precipitation phases at any of the sites. Average

concentrations of S-HCH measured by IADN (Buehler

Page 10: Atmospheric concentrations and deposition of organochlorine pesticides in the US Mid-Atlantic region

ARTICLE IN PRESSR. Gioia et al. / Atmospheric Environment 39 (2005) 2309–23222318

et al., 2001) (Table 4) were similar to levels reported

here. S-HCH concentrations were lower at Jersey City

and Sandy Hook than at the more remote IADN sites,

but this may be an artifact of the breakthrough of a-HCH, which lowered measured gas-phase concentra-

tions in the present study.

Technical HCH mixture typically has a ratio of with

a-HCH/g-HCH ranging from 4 to 7 (Karlsson et al.,

2000). The average a-HCH/g-HCH ratio ranges from

1.371 in Camden to 2.771.8 at the coastal SandyHook. These values are in the same range of those

reported for the NE Atlantic (0.3–4.6) and also for the

Artic (0.9–4.7) (Kelly et al., 1994; Schreitmuller and

Ballschmiter, 1995), indicating that fresh sources of

lindane in New Jersey are unlikely. There is a consistent

seasonal pattern in the a-HCH/g-HCH across New

Jersey. The ratio is o1 during the summer months andincreases to 41 during the winter. The higher concen-tration of lindane observed during the summer probably

arises from differences in the enthalpy of air–surface

exchange between the two compounds (see section on

‘‘Seasonal trends and Relationship with temperature’’

below) or from current usage of lindane. Since g-HCHhas a higher vapor pressure than that of a-HCH, highatmospheric temperature values may accelerate the

volatilization of lindane from surfaces contributing to

higher g-HCH gas-phase concentration in summer.

Because HCHs were not detected in the particle and

precipitation phases, their wet and dry particle deposi-

tion fluxes were not calculated and assumed to be

negligible. As with S-chlordane, S-DDT and aldrin/

dieldrin, the gas absorption was the highest at Camden

and at New Brunswick (42 and 43 ngm�2 day�1,

respectively) as well as the gas-phase atmospheric

concentrations. Again, due to breakthrough of alpha,

the deposition calculations are minimum estimates.

3.6. Endosulfan I and II

Endosulfan is a chlorinated hydrocarbon insecticide

and acaricide of the cyclodiene subgroup, which acts as

a poison to a wide variety of insects and mites on

contact. Endosulfan is a highly toxic pesticide in EPA

toxicity class I. It is a restricted use pesticide (Rama-

narayan and Allen, 2000).

Endosulfan I is the most abundant component in the

mixture and its concentration is always higher than

endosulfan II over all sites in New Jersey. As with many

of the other pesticides, gas-phase concentrations of

endosulfan I were generally highest at Camden and New

Brunswick (Fig. 2a, Table 1). Both endosulfans were

detected in the particle phase at some sites (Table 2),

with the particle phase typically comprising 2–8% and

17–70% of the total atmospheric burdens of endosulfan

I and II, respectively. Both endosulfans were detected in

precipitation (Table 3).

Gas absorption constitutes 490% of the total atmo-

spheric deposition of endosulfan I. Dry and wet

deposition constitute a larger fraction of the total

deposition for endosulfan II because 17–70% of all

atmospheric endosulfan II measured was in the particle

phase, vs. o8% for endosulfan I.

3.7. Dependence of gas-phase OCP concentrations on

temperature

Gas-phase concentrations of most OCPs in New

Jersey were generally higher in summer than winter ,with

the exception of HEPT, which shows a relatively

constant concentration throughout the year. DDT

concentrations were higher between April (mid-spring)

and October (mid-fall) at all sites, with a large

spike of o,p0-DDT in June at the urban Jersey

City site.

The Clausius–Clapeyron equation describes the rela-

tionship between ambient temperature and the gas-

phase partial pressure of semi-volatile organic com-

pounds (Hoff et al., 1992b; Wania et al., 1998)

ln p ¼ �DHenv

RTþ c, (6)

where p is the partial pressure of the compound (Pa),

Henv is a characteristic environmental phase-transition

energy of the compound (kJmol�1), R is the gas

constant (8.314 Pam3mol�1 ¼ kJmol�1) and T is the

temperature (Kelvin). The Clausius–Clapeyron relation-

ship can be expressed graphically as a plot of ln p vs.

1=T : A negative slope was found in all cases where theregression was statistically significant. Aldrin displays

no significant correlation between the gas-phase con-

centration and the 1=T at any of the sites, possibly due

to the small number of samples in which Aldrin was

detected. The negative slopes found for all other

pesticides indicate that the gas-phase concentration

increased with increasing temperature. An increase in

temperature can also affect other environmental

processes that can yield an increase or decrease in

atmospheric gas-phase concentrations of pesticides

such as air/water exchange and volatilization from

terrestrial surfaces. As a result, Henv is useful to

understand the environmental behavior of these com-

pounds. In Table 6Henv for each OCP is compared to its

enthalpies of vaporization (Hvap) and octanol–air

partitioning (Hoa).

All Henv are lower than Hvap and Hoa, which means

that these compounds volatilize more easily from soil

than pure octanol or the pure liquid itself. Gas-phase

concentration of TC, CC, TN, CN, oxychlordane and

HEPTX were significantly correlated with 1=T

(Po0:001) among all sites in New Jersey, with the

exception of Sandy Hook, with the r2 values ranging

from 0.68 to 0.7 in Camden, from 0.3 to 0.6 in Jersey

Page 11: Atmospheric concentrations and deposition of organochlorine pesticides in the US Mid-Atlantic region

ARTICLE IN PRESS

Table 6

Phase transition energies (DHenv795% confidence limits) in kJmol�1 vs. DHvap in kJmol�1 from Hinckley et al. (1990) and DHoa in

kJmol�1 from Shoeib and Harner (2002)

Compound Delaware Bay

DHenv

Camden

DHenv

Jersey City

DHenvNew Brunswick

DHenv

Pinelands

DHenv

Sandy Hook

DHenv DHvap DHoa

a-HCH 23712 2977 3879 1974 2675 2079 68.5 61.8

g-HCH 60722 8679 57712 4779 7376 49714 70.5 65.3

HEPT 58720 1676 NS NS 1675 NS 76.5 66.1

HEPTX 38715 8379 112723 5579 3779 NS

Oxychlordane 44717 6878 58721 8277 5379 NS

TC 49717 6078 48711 47710 5979 41713 80.7 96.3

CC 54716 6577 61711 6078 6478 44714 82.0 98.0

TN 56716 6878 61711 6678 68710 47714 85.5 105.2

CN 108748 86710 58719 6576 57712 43717 83.8

Endosulfan I 132734 175712 127721 136712 112722 89725 80.4

Endosulfan II 158716 66722 94710 51734 86730 82.4

4,40-DDE 92719 6277 3576 7279 6679 58717 87.2 97.8

2,40-DDE 51730 7176 3378 5378 NS NS

4,40 DDT 142758 73714 49725 7379 NS NS 93.2 88.0

2,40 DDT 6277 76727 35715 NS NS 88.6 87.8

Dieldrin 68722 68712 116732 68716 64715 49722 82.5 72.5

NS ¼ not statistically significant.

R. Gioia et al. / Atmospheric Environment 39 (2005) 2309–2322 2319

City, from 0.5 to 0.86 in New Brunswick and from 0.42

to 0.6 in Pinelands. Slopes are highest at Camden and

lowest at Sandy Hook (Table 6). The average Henv

values determined for TC, CC, TN and CN are

significantly lower (Po0:001) at the Sandy Hook siteas compared with the land locked urban and suburban

locations. Sandy Hook is coastal marine site and the

presence of the large water mass may decrease the

influence of temperature on the S-chlordane concentra-tions (Cortes et al., 1998). The ocean water can act as a

sink or a source for these compounds and, in general,

temperature has a smaller effect on the volatilization

and deposition from water bodies than on volatilization

from terrestrial surfaces (Cortes et al., 1998, 1999). The

average Henv at the Sandy Hook for S-chlordanes issimilar to those measured for S-chlordane on the GreatLakes in 1994 (48–44 kJmol�1) (Cortes et al., 1998).

Whereas, Henv measured at Camden and New Bruns-

wick are in the same range of those measured in

Michigan and Arkansas (73 and 71 kJmol�1) (Hoh and

Hites, 2004). HEPT is the only chlordane that does not

display a seasonal pattern. Other investigators have

similarly noted no correlation between the gas-phase

HEPT concentration and 1=T including Hoff et al.

(1992a ,b) in southern Ontario and Janunten et al. (2000)

in Alabama. Table 6 indicates that gas-phase concentra-

tions of g-HCH display stronger temperature dependen-cies (Po0:001) than those of a-HCH. The differences inslopes suggest that a-HCH levels may be governed

primarily by transport, whereas the g-HCH levels are

driven by temperature changes via evaporation from

nearby terrestrial surfaces. Many studies have shown

that a steep slope indicates that the air concentrations

are controlled by re-volatilization from surfaces, while a

flatter slope indicates that advection of air is governing

atmospheric concentration levels (Bidleman et al., 1999;

Cope et al., 1995; Cortes et al., 1999; Yeo et al., 2003).

However, increased breakthrough during warmer peri-

ods may be partially responsible for the steeper slope for

a-HCH.Correlations between ln p vs. 1=T for the DDTs were

statistically significant at Camden, Jersey City and New

Brunswick (Po0:001). Henv estimated for DDTs at

Camden and Jersey City were statistically similar (6277to 73714 kJmol�1). Results of regressions for the NewBrunswick location show a flatter slope than the

Camden site with Henv ranging from 35715 to

7379 kJmol�1. Values found for Henv in Camden are

higher than those estimated in the Great lakes for p,p0-

DDE and p,p0-DDT (between 30 and 57 kJmol�1)

(Cortes et al., 1998), whereas Henv values estimated at

the New Brunswick location are in that range. Correla-

tion with temperature for p,p0-DDE was statistically

significant over all sites in New Jersey (Po0:001). Thestrong relationship between temperature and gas-phase

concentration and the higher Henv in Camden and New

Brunswick suggests a partitioning between the surfaces,

and this can be related to the generally higher

OCP concentrations at Camden and New Brunswick,

suggesting that these two sites are near the atmospheric

Page 12: Atmospheric concentrations and deposition of organochlorine pesticides in the US Mid-Atlantic region

ARTICLE IN PRESSR. Gioia et al. / Atmospheric Environment 39 (2005) 2309–23222320

sources of OCPs, whereas a weak correlation with

temperature for DDTs among the other sites is

indicative of processes that have an opposite weak

correlation with temperature such as long-range trans-

port and direction of local winds. Endosulfan I and II

show the highest r2 and highest Henv values of all the

OCPs.

4. Conclusions

Most of the pesticides measured in this work display

their highest concentrations at the urban/industrial

centers of Camden and New Brunswick, although in

many cases the differences in geometric mean concen-

trations are not statistically significant. High atmo-

spheric OCP levels are likely to be associated with past

or present pesticide use (i.e., agricultural areas). The

data in this study reveal no significant correlation

between urbanization and pesticide concentrations.

However, this study involves only six sites; data from

additional locations and data from chiral analysis of

some of these pesticides would be necessary to confirm

this conclusion.

If so, OCP concentrations would be significantly

higher at Jersey City. The absence of a correlation

between urbanization and high atmospheric OCP levels

suggests that urban sources such as landfills and

brownfields are not important sources of atmospheric

OCPs. Were landfills significant sources of OCPs to the

atmosphere, the presence of more than 10 current or

former landfills in the New Jersey meadowlands would

almost certainly lead to high OCP concentrations at

Jersey City. However, additional data from chiral

analyses of these pesticides should be obtained to clarify

the source. These data represent the first measurements

of OCPs in New Jersey air, and has allowed the first

estimates of atmospheric deposition of such pesticides to

the important coastal zone of the New York–New Jersey

Harbor and Bight. Seasonal variations in gas-phase

concentrations of OCPs can be attributed to ambient

temperature fluctuations. Significant temperature (i.e.,

1=T) dependencies were found for all OCPs at every site,

with the exception of aldrin and HEPT (Table 6).

Acknowledgments

The authors wish to acknowledge T. Glenn IV, D.

Van Ry, I. Koelliker and D.Visco for assistance in

laboratory and fieldwork. This research was supported

by the New Jersey Department of Environmental

Protection and New Jersey Agricultural Experiment

Station of Rutgers University.

References

Alegria, H.A., Bidleman, T.F., Shaw, T.J., 2000. Organochlor-

ine pesticides in ambient air of Belize, Central America.

Environmental Science and Technology 34, 1953–1958.

Bamford, H.A., Offenberg, J.H., Larsen, R.K., Ko, F.C.,

Baker, J.E., 1999. Diffusive exchange of polycyclic aromatic

hydrocarbons across the air–water interface of the Patapsco

River, an urbanized subestuary of the Chesapeake Bay.

Environmental Science and Technology 33, 2138–2144.

Bamford, H.A., Ko, F.C., Baker, J.E., 2001. Seasonal and

annual air–water exchange of polychlorinated biphenyls

across Baltimore Harbor and the Northern Chesapeake

Bay. Environmental Science and Technology 36, 4245–4252.

Bidleman, T.F., Alegria, H., Ngabe, B., Green, C., 1998a.

Trends of chlordane and toxaphene in ambient air of

Columbia, South Carolina. Atmospheric Environment 32,

1849–1856.

Bidleman, T.F., Janunten, M.M., Wiberg, K., Harner, T.,

Brice, K.A., Su, K., Falconer, R.L., Leone, A.D., Aigner,

E.J., Parkhurst, W.J., 1998b. Soil as a source of atmospheric

heptachlor epoxide. Environmental Science and Technology

32, 1546–1548.

Bidleman, T.F., Patton, G.W., Hinckley, D.A., Walla, M.D.,

Cotham, W.E., Hargrave, B.T., 1999. Chlorinated pesticides

and polychlorinated biphenyls in the atmosphere of the

Canadian Artic. In: Kurtz, D. (Ed.), Long-Range Transport

of Pesticides. Lewis Publisher, Chelsea, MI.

Bidleman, T.F., Jantunen, L.M.M., Helm, P.A., Brorstrom-

Lunden, E., Juntto, S., 2002. Chlordane enantiomers and

temporal trends of chlordane isomers in Arctic air.

Environmental Science and Technology 36, 539–544.

Brunciak, P.A., Dachs, J., Gigliotti, C.L., Nelson, E.D.,

Eisenreich, S.J., 2001. Atmospheric polychlorinated biphe-

nyl concentrations and apparent degradation in coastal

New Jersey. Atmospheric Environment 35, 3325–3339.

Buehler, S.S., Basu, I., Hites, R.A.A., 2001. Comparison of

PAH, PCB, and pesticide concentrations in air at two rural

sites on Lake Superior. Environmental Science and Tech-

nology 35, 2417–2422.

Cope, R.H., Manning, M.A., Foremann, W.T., Goolsby, D.A.,

Majewski, M.S., 1995. Occurrence of pesticides in rain and

air in urban and agricultural area of Mississippi, April–-

September. The Science of the Total Environment 248,

227–240.

Cortes, D.R., Basu, I., Sweet, C.W., Brice, K.A., Hoff, R.M.,

Hites, R.A., 1998. Temporal trend in gas-phase concentra-

tions of chlorinated pesticides measured at the shores of the

Great Lakes. Environmental Science and Technology 32,

1920–1927.

Cortes, D.R., Hoff, R.M., Brice, K.A., Hites, R.A., 1999.

Evidence of current pesticide use from temporal and

Clausius–Clapeyron plots: a case study from integrated

atmospheric deposition network. Environmental Science

and Technology 33, 2145–2150.

Duerksen-Hughen, P., Patricia, R.-T., Sanjivani, D., Robin, L.,

1993. US Department of Health and Human Services.

Public Health Service, Agency for Toxic Substances and

Disease Registry April.

Eitzer, B.D., Iannucci-Berger, W., Mattina, I.M., 2003.

Volatilization of weathered chiral and achiral chlordane

Page 13: Atmospheric concentrations and deposition of organochlorine pesticides in the US Mid-Atlantic region

ARTICLE IN PRESSR. Gioia et al. / Atmospheric Environment 39 (2005) 2309–2322 2321

residues from soil. Environmental Science and Technology

37, 4887–4893.

Faroon, O., Olivia, H., Fernando, L., Steven, S., Gloria, S.,

Mario, C., Daniell, G., 2002. US Department of Health and

Human Services. Public Health Service, Agency for Toxic

Substances and Disease Registry September.

Franz, T.P., Eisenreich, S.J., Holsen, T.M., 1998. Dry deposi-

tion of particulate polychlorinated biphenyls and polycyclic

aromatic hydrocarbons to Lake Michigan. Environmental

Science and Technology 32, 3681–3688.

Gigliotti, C.L., Dachs, J., Nelson, E.D., Brunciak, P.A.,

Eisenreich, S.J., 2000. Polycyclic aromatic hydrocarbons in

the New Jersey coastal atmosphere. Environmental Science

and Technology 34, 3547–3554.

Gigliotti, C.L., Brunciak, P.A., Dachs, J., Glenn, T.R., Nelson,

E.D., Totten, L.A., Eisenreich, S.J., 2002. Air–water

exchange of polycyclic aromatic hydrocarbons in the New

York–New Jersey, USA, Harbor Estuary. Environmental

Toxicology and Chemistry 21 (2), 235–244.

Gioia, R., 2004. Organic carbon and elemental carbon

associated with the PM2.5 mass and organochlorine

pesticides in New Jersey air. Thesis submitted for the degree

of Master of Science at Rutgers University, New Jersey, US.

Halsal, C.J., Bailey, R., Stern, G.A., Barrie, L.A., Fellin, P.,

Muir, D.C.G., Rosenberg, B., Rovinsky, F.Ya., Kononov,

E.Ya., Pastukhov, B., 1998. Multi-year observations of

organohalogen pesticides in the Arctic atmosphere. Envir-

onmental Pollution 102, 51–62.

Hanley, D., Beblo, D.A., David, W., Armando, A., Stephen, B.,

Daniel, P., 2002. US Department of Health and Human

Services. Public Health Service, Agency for Toxic Sub-

stances and Disease Registry September.

Harner, T., Wideman, J.L., Jantunen, L.M.M., Bidleman, T.F.,

Parkhurst, M.J., 1999. Residues of organochlorine pesti-

cides in Alabama soils. Environmental Pollution 106,

323–332.

Harner, T., Shoeib, M., Diamond, M., Stern, G., Rosenberg,

B., 2004. Using passive air samplers to assess urban-rural

trends for persistent organic pollutants. 1. Polychlorinated

biphenyls and organochlorine pesticides. Environmental

Science and Technology 38, 4474–4483.

Henry, G.A., Ronald, B., Paul, F.G., 1994. US Department of

Health and Human Services. Public Health Service, Agency

for Toxic Substances and Disease Registry.

Hillery, B.R., Simcik, M.F., Basu, I., Hoff, R.M., Strachan,

W.J.M., Burniston, D., Chan, C.H., Brice, K.A., Sweet,

C.W., Hites, R.A., 1998. Atmospheric depeosition of toxic

pollutants to the Great Lakes as measured by integrated

atmospheric deposition network. Environmental Science

and Technology 32, 2216–2221.

Hinckley, D.A., Bidlemann, T.F., Foreman, W.T., Tuschall,

J.R., 1990. Determination of vapor pressures for nonpolar

and semipolar organic compounds from gas chromato-

grahic retention data. Journal of Chemical Engineering

Data 35, 232–237.

Hoff, R.M., Muir, D.C.G., Grift, N.P., 1992a. Annual cycle of

polychlorinated biphenyls and organohalogen pesticides in

air in southern Ontario. 1. Air concentration data.

Environmental Science and Technology 26, 266–275.

Hoff, R.M., Muir, D.C.G., Grift, N.P., 1992b. Annual cycle of

polychlorinated biphenyls and organohalogen pesticides in

air in southern Ontario. 2. Atmospheric transport and

sources. Environmental Science and Technology 26,

276–283.

Hoh, E., Hites, A.R., 2004. Sources of toxaphene and other

organochlorine pesticides in North America as determined

by air measurements and potential sources contribution

function analysis. Environmental Science and Technology

38, 4187–4194.

Hung, H., Halsall, C.J., Blanchard, P., Li, H.H., Fellin, P.,

Stern, G., Rosenberg, B., 2002. Temporal trend of organo-

chlorine pesticides in the Canadian Artic atmosphere.

Environmental Science and Technology 36, 862–868.

Iwata, H., Tanabe, S., Sakai, N., Tatsukawa, R., 1993.

Distribution of persistent organochlorines in the oceanic

air and surface seawater and the role of ocean on their

global transport and fate. Environmental Science and

Technology 27, 1080–1098.

Iwata, H., Tanabe, S., Ueda, K., Tatsukawa, R., 1995.

Persistent organochlorine residues in air, water, sediments,

and soils from the Lake Baikal Region, Russia. Environ-

mental Science and Technology 29, 792–801.

Janunten, L.M., Bidleman, T.F., Harner, T., Parkhurst, W.J.,

2000. Toxaphene, chlordane and other organochlorine

pesticides in Alabama air. Environmental Science and

Technology 34, 5097–5105.

Kallenborn, R., Oehme, M., Wynn-Williams, D.D., Schlab,

ach, Harris, J., 1998. Ambient air levels and atmospheric

long range transport of persistent organochlorines to Signy

Island, Antarctica. Science of the Total Environment 220,

167–180.

Karlsson, H., Derek, C., Muir, G., Teixiera, C.F., Burniston,

D.A., Strachan, W.M.J., Hecky, R.E., Mwita, J., Bootsma,

H.A., Grift, N.P., Kidd, K.A., Rosenberg, B., 2000.

Persistent chlorinated pesticides in air, water and precipita-

tion from the Lake Malawi Area, Southern Africa.

Environmental Science and Technology 34, 4490–4495.

Kelly, T.J., Mukund, R., Spicer, C.W., Pollack, A.J., 1994.

Concentrations and transformation of hazardous air

pollutants. Environmental Science and Technology 28,

378–387.

Kucklick, J.R., Baker, J.E., 1998. Organochlorines in Lake

Superior’s Food Web. Environmental Science and Technol-

ogy 32, 1192–1198.

Leone, A.D., Amato, S., Falconer, R.L., 2001. Emission of

chiral organochlorine pesticides from agricultural soils in

the Combelt Region of the US. Environmental Science and

Technology 35, 4591–4596.

Mackay, B., Shiu, W.-Y., Ma, K.-C., 1999. Physical–Chemical

Properties and Environmental Fate Handbook. Chapman &

Hall, London.

McConnell, L.L., Kuclick, J.R., Bidleman, T.F., Ivanov, G.P.,

Chernyak, S.M., 1996. Air–water gas exchange of organo-

chlorine X+compounds on Lake Baikal, Russia. Environ-

mental Science and Technology 36, 2975–2983.

Miller, S.M., Green, M.L., DePinto, J.V., Hornbuckle, K.C.,

2001. Results from the Lake Michigan Mass Balance Study:

concentrations and fluxes of atmospheric polychlorinated

biphenyls and trans-nonachlor. Environmental Science and

Technology 35, 278–285.

Muir, D.C.G., Ford, C.A., Rosenberg, B., Norstrom, R.J.,

Simon, M., Beland, P., 1996. ‘‘Persistent organochlorines in

Page 14: Atmospheric concentrations and deposition of organochlorine pesticides in the US Mid-Atlantic region

ARTICLE IN PRESSR. Gioia et al. / Atmospheric Environment 39 (2005) 2309–23222322

beluga whales (Delphinapterus leucas) from the St. Lawrence

River estuary—I. Concentrations and patterns of specific

PCBs, chlorinated pesticides and polychlorinated dibenzo-p-

dioxins and dibenzofurans. Environmental Pollution 93,

219–234.

Nelson, E.D., Mc Connel, L.L., Baker, J.E., 1998. Diffusive

exchange of gaseous polycyclic aromatic hydrocarbons and

polychlorinated biphenyls across the air–water interface of

the Chesapeake Bay. Environmental Science and Technol-

ogy 32, 912–919.

NJDEP, 1999. Findings & recommendations for the remedia-

tion of historic pesticide contamination. Final Report,

March.

Offenberg, J.H., Nelson, E.D., Gigliotti, C.L., Eisenreich, S.J.,

2004. Chlordanes in New Jersey Air: 1997–1999. Environ-

mental Science and Technology 38, 3488–3497.

Pirrone, N., Keeler, G.J., Holsen, T.M., 1995. Dry deposition

of semivolatile organic compounds to Lake Michigan.

Environmental Science and Technology 29, 2123–2132.

Pirrone, N., Keeler, G.J., Holsen, T.M., 1995. Dry deposition

of trace elements to Lake Michigan: a hybrid-receptor

deposition modeling approach. Environmental Science and

Technology 29, 2112–2122.

Ramanarayan, T., Allen, R., 2000. EPA registration report,

July 2000.

Sahsuvar, L., Helm, P.A., Jantunen, L.M., Bidleman, T.F.,

2003. Henry’s Law Constants for alpha-, beta-, and gamma-

hexachlorocyclohexanes (HCHs) as a function of tempera-

ture and revised estimates of gas exchange in Arctic regions.

Atmospheric Environment 37, 983–992.

Schreitmuller, J., Ballschmiter, K., 1995. Air–water equilibrium

of hexachlorocyclohexanes and chloromethoxybenzenes in

the North and South Atlantic. Environmental Science and

Technology 29, 207–215.

Shoeib, M., Harner, T., 2002. Using measured octanol–air

partition coefficients to explain environmental partitioning

of organochlorine pesticides. Environmental Toxicology

and Chemistry 21, 984–990.

Suntio, L.R., Shiu, W.Y., MacKay, D., Seiber, J.N., Glotfelty,

D., 1988. Critical review of Henry’s Law Constants for

pesticides. Reviews of Environmental Contamination And

Toxicology 103, 1 Tharacad Ramanaraynan, Richard Allen;

EPA Registration Report, July 2000.

Totten, L.A., Gigliotti, C.L., Van Ry, D.A., Offenberg, J.H.,

Nelson, E.D., Dachs, J., Reinfelder, J.R., Eisenreich, S.J.,

2004. Atmospheric concentrations and deposition of poly-

chorinated biphenyls to the Hudson River Estuary.

Environmental Science and Technology 38, 2568–2573.

Van Ry, D.A., Gigliotti, C.L., Glenn, T.R., Nelson, E.D.,

Totten, L.A., Eisenreich, S.J., 2002. Wet deposition of

polychlorinated biphenyls in urban and background areas

of the Mid-Atlantic states. Environmental Science and

Technology 36, 3201–3209.

Walker, K., Vallero, D.A., Lewis, R.G., 1999. Factors

influencing the distribution of lindane and other hexachlor-

ocyclohexanes in the environment. Environmental Science

and Technology 33, 4373–4378.

Wania, F., Haugen, J.-E., Lei, Y.D., Mackay, D., 1998.

Temperature dependence of atmospheric concentration of

semi-volatile organic compounds. Critical Reviews In

Environmental Science and Technology 32 (8).

Withman, W.G., 1923. The two film theory of gas-absorption.

Chemical and Metallurgical Engineering 29, 146–148.

Yeo, H.-G., Choi, M., Chun, M.-Y., Sunwoo, Y., 2003.

Concentration distribution of polychlorinated biphenyls

and organochlorine pesticides and their relationship with

temperature in rural air of Korea. Atmospheric Environ-

ment 37, 3831–3839.

Further Reading

Glassmeyer, S.T., Brice, K.A., Hites, R.A.J., 1999. Atmo-

spheric concentrations of toxaphene on the coast of the

Lake Superior. Journal of Great Lakes Research 25 (3),

492–499.

Hargrave, B.T., Barrie, L.A., Bidleman, T.F., Welch, H.E.,

1997. Seasonality in exchange of organochlorines between

Arctic air and seawater. Environmental Science and

Technology 31, 3258–3266.

Mader, B.T., Pankow, J., 2001. Gas/solid partitioning of

semivolatile organic compounds (SOCs) to air filters. 3.

An analysis of gas adsorption artifacts in measurements of

atmospheric SOCs and organic carbon (OC) when using

Teflon membrane filters and quartz fiber filters. Environ-

mental Science and Technology 35 (17), 3422–3432.