chapter 4: deposition of nitric acid and reactive nitrogen at

32
Chapter 4: Deposition of Nitric Acid and Reactive Nitrogen at Harvard Forest Abstract The hourly gas-phase HNO 3 concentration was measured at the Harvard Forest Environmental Measurement Site during the spring, summer, and fall of 2000 using a Tunable Diode Laser Absorption Spectrometer (TDLAS). A dry deposition inferential method (DDIM) employed micrometeorological data to estimate the deposition velocity of HNO 3 on an hourly basis. HNO 3 fluxes were calculated hourly as the product of measured concentration and inferred deposition velocity. We aggregated the hourly data on daily and weekly timescales to simulate standard dry deposition monitoring procedures which use long-time averaging techniques, such as filterpacks and denuders, to measure concentrations. We found that the longer averages introduced noise, but not significant bias, into the resulting HNO 3 fluxes. HNO 3 concentrations were slightly lower under clean (NW flow) conditions compared to more polluted (SW); inferred V d (HNO 3 ) was nearly the same for the two flow regimes. V d (PAN) was estimated by comparison with V d (O 3 ) on nights were first-order decays of both species were observed and deposition to the surface was likely to be the primary loss mechanism. Using estimated PAN and NO x deposition velocities and measured concentrations, we find that the concentration and flux budgets of NO y are essentially closed for the NW, and unaccounted by up to 50% for the SW. We calculated that the deposition velocity of the unmeasured species (SW flow) was approximately 3 cm s -1 . 104

Upload: others

Post on 04-Feb-2022

2 views

Category:

Documents


0 download

TRANSCRIPT

Chapter 4: Deposition of Nitric Acid and Reactive Nitrogen at Harvard Forest

Abstract

The hourly gas-phase HNO3 concentration was measured at the Harvard Forest

Environmental Measurement Site during the spring, summer, and fall of 2000 using a

Tunable Diode Laser Absorption Spectrometer (TDLAS). A dry deposition inferential

method (DDIM) employed micrometeorological data to estimate the deposition velocity

of HNO3 on an hourly basis. HNO3 fluxes were calculated hourly as the product of

measured concentration and inferred deposition velocity. We aggregated the hourly data

on daily and weekly timescales to simulate standard dry deposition monitoring

procedures which use long-time averaging techniques, such as filterpacks and denuders,

to measure concentrations. We found that the longer averages introduced noise, but not

significant bias, into the resulting HNO3 fluxes. HNO3 concentrations were slightly

lower under clean (NW flow) conditions compared to more polluted (SW); inferred

Vd(HNO3) was nearly the same for the two flow regimes. Vd(PAN) was estimated by

comparison with Vd(O3) on nights were first-order decays of both species were observed

and deposition to the surface was likely to be the primary loss mechanism. Using

estimated PAN and NOx deposition velocities and measured concentrations, we find that

the concentration and flux budgets of NOy are essentially closed for the NW, and

unaccounted by up to 50% for the SW. We calculated that the deposition velocity of the

unmeasured species (SW flow) was approximately 3 cm s-1.

104

4.1. Introduction

Reactive nitrogen trace gas species play an important role in tropospheric

photochemistry by stimulating ozone formation, affecting the concentration of hydroxyl

radicals, and contributing to acid precipitation. Emissions of reactive nitrogen are largely

in the form of nitric oxide, NO, predominantly from combustion of fossil fuel and

production by soil microbes. Additional sources of NO to the troposphere come from

biomass burning, lightning, ammonia oxidation, emission from the oceans and transport

from the stratosphere. NO is quickly oxidized to nitrogen dioxide, NO2, by reaction with

ozone, O3, or peroxy radical species, RO2. During the day, NO2 may be photolyzed back

to NO, and the product oxygen atom regenerates O3 by combining with molecular O2.

When NO oxidation proceeds via RO2 instead of O3, NO2 photolysis results in net

production of O3.

NOx is converted to nitric acid, HNO3, via both homogeneous and heterogeneous

chemical pathways. Recycling of HNO3 back to NOx is very inefficient in the

troposphere compared to wet and dry deposition, and thus oxidation of NOx to HNO3

represents a termination step for NOx photochemistry in the troposphere. NOx may also

form temporary non-radical reservoir species such as peroxyacetyl nitrate (PAN) and

other organic nitrates, allowing alternative deposition pathways or export to remote

regions where radicals may be regenerated. Thus conversion of NOx radicals to HNO3

and subsequent deposition is considered the primary loss process for reactive nitrogen

and an important sink for HOx radicals, which participate in the oxidation of

hydrocarbons to form ozone [Logan et al., 1981; Logan, 1983].

105

Deposition of HNO3 contributes acid to ecosystems second only to sulfuric acid

deposition in the northeastern United States [U.S. Environmental Protection Agency,

2001]. HNO3 can act as a fertilizer, delivering fixed nitrogen to forests and crops, or as a

damaging toxicant, stripping base cations and mobilizing toxic elements such as

aluminum in soils. Thus the partitioning of total reactive nitrogen, NOy = NO + NO2 +

NO3 + N2O5 + HNO3 + peroxyacetylnitrate (PAN) + other organic nitrates + aerosol

nitrate, among the individual species, has direct effects on photochemical production of

ozone in rural and urban areas, on deposition of acidic species to ecosystems, and on the

export of atmospheric nitrogen from the continent to the global troposphere.

Comparisons have been made previously between the sum of concentrations of

individual NOy species, Σ(NOy)i, and total NOy measured by catalytic conversion (on hot

gold or molybdenum) to NO at continental rural sites [e.g. Parrish et al., 1993; Aneja et

al., 1996; Thornberry et al., 2001], coastal sites [e.g. Harrison et al., 1999], and for the

remote free troposphere [e.g. Bradshaw et al., 2000]. Discrepancies between [NOy] and

Σ(NOy)i have been largely attributed to interference of reduced nitrogen species such as

NH3, HCN, and other organic molecules with direct C-N bonds, in the NOy

measurements [Kliner et al., 1997; Bradshaw et al. 1998; Harrison et al., 1999]. Taking

into account possible interference, total NOy measurements appear to be dominated by

NOx, HNO3, and PANs, with lesser contributions from aerosol nitrates, under a wide

variety of lower tropospheric conditions.

Few studies have attempted to speciate the deposition flux of reactive nitrogen to

the surface. Munger et al. [1996; 1998] observed concentrations and eddy covariance

fluxes of NOy and NOx, and inferred that nearly all dry deposition of NOy occurs in the

106

form of rapidly-depositing species such as HNO3, with a summertime contribution due to

organic hydroxyalkyl nitrates formed during the oxidation of biogenic hydrocarbons

(isoprene and monoterpenes). We have confirmed (see Chapter 3) that direct deposition

of NO and NO2 are small compared to NOy fluxes. Prior estimates of PAN deposition

velocity to vegetation also indicate that it should contribute negligibly to the flux of NOy

[Shepson et al., 1992; Schrimpf et al., 1996; McFayden and Cape, 1999].

Due in part to its tendency to adsorb onto many types of surfaces, HNO3

measurement techniques with demonstrated accuracy, high precision, and a lack of

interferences have taken decades to develop. Increasingly reliable methods have

appeared over the last several years. Filter packs and denuders of various designs are

employed for routine monitoring [e.g. Harrison et al., 1999; Rosman et al., 2001; Clarke

et al., 1997]. These techniques often require collection times on the order of hours to

days, after which the filter or denuder is removed, extracted, and processed to detect NO3-

in solution by ion chromatography. Potential interferences inherent to filters and

denuders have been documented and minimized during the last two decades.

Nevertheless, positive and negative artifacts remain possible under certain conditions,

and the degree to which various organic nitrate-containing species are sampled remains

unknown. Mist chamber techniques rely on capture and dissolution of HNO3 and

detection by ion chromatography; recent implementations have the advantage of much

shorter integration times (minutes) and thorough calibration procedures, which have

greatly reduced the uncertainties associated with the measurements [cf. Talbot et al.,

1997; Bradshaw et al., 1998]. Day et al. [2002] employ thermal dissociation of NOy

compounds at calibrated temperatures (including HNO3), followed by laser-induced

107

fluorescence detection of NO2. This instrument design has the advantage of measuring

peroxy nitrates and alkyl nitrates in addition to HNO3.

Recently developed Chemical Ionization Mass Spectrometers (CIMS) have

demonstrated faster time response (~1 s) for HNO3 measurements. Sensitivity in these

instruments is also very good, under 50 pmol mol-1 HNO3 concentration in air, and

interferences have been minimized through careful selection of ion chemistry and

standard addition calibrations of HNO3 from heated, constant-flow permeation tubes

[Huey et al., 1998; Mauldin et al., 1998; Miller et al., 2000; Furutani and Akimoto,

2002]. Permeation devices require long stabilization times, and very constant

temperatures and flow rates must be maintained in all parts of the calibration oven and

delivery tubing to ensure quantitative HNO3 calibrations. Nevertheless, recently

developed systems have agreed well in intercomparisons and have no obvious or large

interferences [e.g. Fehsenfeld et al., 1998].

Tunable Diode Laser Absorption Spectroscopy (TDLAS) provides a useful

alternative for continuous, unattended field monitoring of HNO3. The TDLAS

instrument used in this study employs recent advances in long path length/low volume

absorption cells, inlet material and design, and real-time data processing to achieve a

spectroscopically-calibrated measurement of HNO3 concentration. A prototype HNO3

TDLAS instrument participated in an informal intercomparison with the CIMS HNO3

instrument described by Huey et al. [1998] in 1996. The TDLAS compared well with the

CIMS results over a wide range of concentrations, ambient relative humidity, and

pollution levels [see Chapter 2; Horii et al., 1999].

108

We present measurements of NOy and HNO3, along with NO, NO2, and PAN

during summer and fall of 2000 at the Harvard Forest Environmental Measurement Site.

In order to disaggregate the flux of reactive nitrogen by species, we measured the eddy

covariance flux of NOy, inferred the flux of HNO3 using a standard approach, and

estimated the contributions of NOx and PAN to NOy deposition. Using the speciated NOy

concentrations and fluxes, we determined the concentration and average deposition

velocity of a class of rapidly-depositing compounds present under the more polluted,

warmer, and more humid flow regime at the site.

4.2. Methods

4.2.1. Experimental Methods

HNO3 and NOy were measured at the Harvard Forest Environmental

Measurement Site in central Massachusetts, along with micrometeorological and

supporting data, during the summer and fall of 2000 (the site is described more fully in

Chapter 3). The concentration of HNO3 was measured using a tunable diode laser

absorption spectrometer (TDLAS) installed on a tower above the forest canopy (refer to

Chapter 2 for full description of the instrument). The inlet was designed to keep the

residence time short, to minimize wall effects, and to exclude aerosols from the sample

flow. Sampling height for HNO3 was 22 m, ~2 m above the canopy top. The TDLAS

specifically and quantitatively measured gas-phase HNO3, using spectroscopic fitting of

several absorption lines to eliminate interferences from other reactive nitrogen species

such as PAN, hydroxyalkyl nitrates, NOx, and fine aerosols. These species are however

all included in the NOy measurement. Ongoing NOy concentrations and eddy covariance

109

flux measurements on a nearby tower at a sampling height of 29 m employed a hot gold

catalyst to reduce total nitrogen oxide to NO in the presence of H2. The catalyst was

located at the inlet so that species with the tendency to adsorb onto tubing walls, such as

HNO3, were not lost. A complete description of the NOy instrument can be found in

Munger et al. [1996]. PAN was measured using capillary-column gas chromatography

(Metcon Inc.) with electron-capture detection. The instrument calibration was checked

periodically by introduction of PAN generated by photolysis of acetone in the presence of

NO2 [Warneck and Zerbach, 1992].

4.2.2. Data Analysis Methods

Dry Deposition Inferential Method for HNO3

The hourly deposition velocity of HNO3 was estimated using a dry deposition

inferential method (DDIM) in which the deposition velocity, Vd, is modeled as a set of

resistances in series depending on meteorological and site-specific conditions. The flux

was computed as the product of measured HNO3 concentration and inferred Vd.

[HNO3] Vd = [HNO3] (Ra + Rb + Rc)-1 (Eq. 4.1)

Ra represents aerodynamic resistance, Rb approximates the diffusive boundary layer

resistance, and Rc is a species-specific resistance to uptake. High solubility and surface

reactivity make Rc ≈ 0 for HNO3 [Wesely and Hicks, 2000; Hanson and Lindberg, 1991].

After Wesley and Hicks [1977], Hicks et al. [1987], Wesley [1989], and Meyers et al.

[1989], Ra and Rb were calculated as follows:

110

**2 kuuuR H

−= (Eq. 4.2)

LdzstableH

)(5)( −−=ψ (Eq. 4.3a)

−−

−−

+=2

)(ln09.0)(ln39.0598.0exp)(L

dzL

dzunstableHψ (Eq. 4.3b)

*1.7

Pr*2 3

2

uSc

kuRb ≅

≅ (Eq. 4.4)

where u is the horizontal wind speed measured by the sonic anemometer; u* is the

friction velocity (equal to the square root of the absolute value of the momentum flux,

also measured by the sonic anemometer); k is von Karman’s constant, 0.4; Ψh is a

diabatic stability correction coefficient for heat transfer; z is the sensor height; d is the

zero-plane displacement height, approximately 19.8 m at Harvard Forest [Lefer et al.,

1997]; Sc is the Schmidt number for HNO3, 1.22; Pr is the Prandtl number for air, 0.72; L

is the Monin-Obukhov length scale:

h

p

QgkTuC

L⋅⋅

⋅⋅⋅−=

3*ρ (Eq. 4.5)

ρ is the density of air; Cp is the heat capacity of air; T is air temperature; and Qh is the

eddy covariance sensible heat flux. Stable conditions are indicated by positive z/L,

unstable by negative z/L.

Because we have made the approximation that Rc is negligible for HNO3, the

inferred deposition velocity depends sensitively on Ra and Rb. Ra in particular is known

to be poorly approximated by Eq. 4.2 under very stable conditions and over uneven

terrain [Brook et al., 1997; Wesely and Hicks, 2000]. However, a better technique for

111

estimating Vd(HNO3) without a multi-layer aerodynamic and boundary layer resistance

model has not been developed.

4.3. Results

4.3.1. HNO3 Concentrations and Inferred Fluxes

In Figure 4.1, the measured HNO3 concentrations and inferred deposition

velocities are shown on a diel basis, separated into the dominant wind sectors of

Northwest (270°-45°) and Southwest (180°-270°) for summer and fall 2000. As in

Chapter 3, -Vd or “exchange velocity” is plotted, so that negative values represent

deposition. At Harvard Forest, northwesterly flows correspond to relatively cool, dry,

and unpolluted conditions, and southwest winds transport warmer, more humid, and

significantly more polluted air masses to the site [Moody et al., 1998]. On average, the

HNO3 concentration has a slight diel cycle in the summer, and is nearly constant over the

day during the fall. All data collected during the spring, summer, and fall of 2000,

segregated by wind direction, are considered together in Figure 4.2. There is remarkably

little difference in the observed HNO3 concentration between Northwest and Southwest

flow regimes in both seasons. The inferred deposition velocity, in contrast, has a strong

diel cycle, driven by the larger u* during the day than at night. A sample time series in

Figure 4.3 shows that the inferred HNO3 deposition flux exceeds the measured downward

NOy flux during some hours, but the two generally follow one another. Conditions when

deposition flux of NOy greatly exceeds the sum of measured, inferred, and estimated

fluxes of NOx, HNO3, and PAN are discussed in section 4.4 below.

112

Hour

9 14 10 88 7

910 11 14 14 17 14

2015 19

17 107 13

13

0 6 12 18

Fall NW

77

46 5

6

4

64

45 3

6

11

12 119

8 66

9

0 6 12 18

Summer NW

[HN

O3]

(nm

ol m

ol-1)

-Vd(

HN

O3)

(cm

s-1)

[HN

O3]

(nm

ol m

ol-1)

-Vd(

HN

O3)

(cm

s-1)

mean ± st. dev.median25th & 75th quantiles

3 43 6 4

34

48

1111

9

11 1210

9

10

95

53

0 6 12 18

Summer SW

Hour

4

6

5

3

89 9

710 13 11

1317

1314 14 12

6

6

610

0 6 12 18

Fall SW

-8-6

-4-2

01

23

-8-6

-4-2

01

23

-8-6

-4-2

01

23

-8-6

-4-2

01

23

Figure 4.1. Mean (open symbols, dashed lines) and median (solid symbols, solid lines)

diel concentrations and inferred deposition velocities of HNO3. The 25th and 75th

quantiles are shown as solid traces above and below the medians. Dashed vertical lines

represent the standard deviations of the means. Number of hours of data are shown

above the HNO3 concentrations.

113

0.0

0.4

0.8

0.0

0.4

0.8

SWNW

All Data

-8-6

-4-2

0

Hour

0 5 10 15 20

-6-4

-20

[HN

O3]

(nm

ol m

ol-1

)V(

HN

O3)

(cm

s-1

)V

·[H

NO

3] (n

mol

mol

-1cm

s-1)

Figure 4.2. Median diel HNO3 concentrations (upper panel), inferred deposition

velocities (center panel), and flux, V(HNO3)·[HNO3], (lower panel) for Southwest and

Northwest wind sectors, summer and fall, 2000.

114

25

20

15

10

5

0

[NO

y], [H

NO

3] (n

mol

mol

-1)

286284282280278276274272270

Day of Year 2000

-25

-20

-15

-10

-5

0

FNO

y, FH

NO

3 (µm

ol m

-2 h

r-1)

NOy HNO3 HNO3 + NOx

Figure 4.3. Time series from 25 September to 12 October 2000. [NOy] and [HNO3] are

in the upper panel, along with [HNO3]+[NOx]. Eddy covariance NOy flux and inferred

HNO3 flux are shown in the lower panel.

115

Several dry deposition measurement networks employ a modified DDIM to

estimate fluxes of SO2, O3, HNO3, and particles [e.g. Clarke et al., 1997; Meyers et al.,

1998]. For HNO3, filterpacks are left in place for a week at a time to measure average

weekly concentrations. In the case of SO2, diel concentration and inferred deposition

velocity correlate strongly, with highest values in the daytime; when average weekly SO2

concentration is used to calculate the flux, FSO2 can be systematically underestimated by

as much as 40% [Matt and Meyers, 1993]. The hourly Harvard Forest dataset provides

an opportunity to test whether a similar bias exists for weekly average HNO3 fluxes.

We defined data weeks for Harvard Forest during the summer and fall of 2000

beginning and ending on Tuesdays at 9:00 AM local time, an arbitrary standard employed

by the Clean Air Status and Trends Network (CASTNet) and others. Data days begin and

end at midnight Eastern Standard Time. For days with more than 5 hours of HNO3

concentration data, and weeks with more than 10 hours of data (20 weeks), we calculated

daily and weekly average (A) and integrated (I) fluxes using the hourly inferred

deposition velocities and weekly average or hourly concentrations according to Equations

(4.6) and (4.7).

weekhourhourhours

HNOVdN

A ][13∑= (Eq. 4.6)

hourhourhourhours

HNOVdN

I ][13∑= (Eq. 4.7)

In none of these 20 weeks did the HNO3 data coverage amount to the full 168 hours.

However, the distribution of data gaps was nearly uniform over all hours of the day and

night. The weekly correlation coefficient of deposition velocity with concentration did

not increase or decrease with the number of data hours in the day or week.

116

Figure 4.4 shows the daily and weekly averaged vs. integrated HNO3 fluxes. In

both cases, the results fall near the 1:1 line. Also shown are the daily and weekly

residuals (A-I) plotted against the quantiles of the standard normal. Residuals appear

near-normally distributed, with extended wings in the daily case. The mean (± 1 standard

deviation) correlation coefficients were r(Vd,HNO3)daily = -0.1±0.5 and r(Vd,HNO3)weekly

= 0.1±0.3, indicating no significant correlation on either timescale (cf. Figure 4.2). For

weeks with 20 or more hours of data, r(Vd, HNO3)weekly was as large as ±0.5, but unlike

Matt and Meyers’ [1993] results for SO2, there was no consistent positive or negative

correlation. Thus it appears that differences between A and I were largely due to random

errors in both deposition velocity and concentration. Diel correlations between HNO3

concentration and Vd are much smaller than for SO2 and did not introduce a substantial

bias in the weekly average flux, compared to weekly integrated flux, over the 20 data

weeks considered here.

4.3.2. Estimates of NOx and PAN Fluxes

In order to evaluate the contributions of HNO3, NOx, and PAN to the total

deposition flux of NOy, we use the inferred hourly flux of HNO3 discussed above, the

parameterization of NO and NO2 fluxes presented in Chapter 3, and an estimate of the

nighttime deposition velocity of peroxyacetyl nitrate (PAN), CH3C(O)OONO2. In order

to obtain an estimate of the deposition part of the NO and NO2 fluxes during both day

and night, we calculate FNOx(dep) by excluding the light-dependent, coupled flux term in

the parameterizations (see Chapter 3 for details):

FNOx(dep) ≈ V0 · [NO2] + a · [NO2]2 + Γ · Vd(O3), (Eq. 4.8)

117

-15 -10 -5 0

-15

-10

-50

Quantiles of Standard Normal

-2 -1 0 1 2

-2-1

01

2

-12 -10 -8 -6 -4 -2 0

-12

-10

-8-6

-4-2

0

Quantiles of Standard Normal

-2 -1 0 1 2

-10

1

Aver

aged

FH

NO

3(µ

mol

m-2

hr-1

)

Daily Weekly

Aver

aged

FH

NO

3(µ

mol

m-2

hr-1

)

Integrated FHNO3 (µmol m-2 hr-1) Integrated FHNO3 (µmol m-2 hr-1)

Res

idua

l (A-

I) FH

NO

3(µ

mol

m-2

hr-1

)

Res

idua

l (A-

I) FH

NO

3(µ

mol

m-2

hr-1

)

1:1 1:1

Figure 4.4. Upper plots: daily (left) and weekly (right) averaged inferred fluxes vs.

integrated inferred fluxes, with 1:1 line. Lower plots: daily (left) and weekly (right)

residual integrated minus averaged HNO3 flux versus quantiles of the Standard Normal.

Residuals appear near normally-distributed.

118

where V0 = -0.21 cm s-1, a = -0.013 nmol-1 mol cm s-1, and Γ = 0.21 (see Chapter 3 for

associated errors). Equations (4.9) may be considered an upper limit to nighttime NOx

deposition, since it is possible that heterogeneous NO2 hydrolysis may produce gas-phase

HONO at night, regenerating NOx and associated redox chemistry the following morning

upon photolysis.

Using selected nights at Harvard Forest when first order decay of [PAN] and [O3]

were observed, we obtained an estimate of the deposition velocity of PAN assuming that

the primary loss mechanism of both species is deposition to the surface. This assumption

is reasonable for O3, where the nighttime chemistry is well characterized, but would

render the estimate of Vd(PAN) an upper limit if unknown loss reactions were important.

As described in detail in Appendix B, we find Vd(PAN) ≤ 0.8 ± 0.4 cm s-1.

4.3.3. NOy Concentration and Flux Budget

Median diel cycles of the measured concentrations of NO, NO2, HNO3, PAN, and

NOy are plotted in Figure 4.5 for June, July, and August 2000, for Northwest and

Southwest wind sectors. The measured flux of NOy is shown along with the inferred flux

of HNO3, the parameterized NO and NO2 deposition fluxes, and estimated PAN flux.

The latter was obtained by multiplying the estimated nighttime deposition velocity, 0.8

cm s-1, with the median diel PAN concentration. Figure 4.6 follows the same format as

Figure 4.5 for September, October, and November, omitting PAN concentration and flux

because the instrument was not operational after 18 July 2000. Table 4.1 summarizes the

fractional contributions of NOx, HNO3, and PAN to total NOy concentration and flux by

wind sector, day and night, and season, obtained by finding the slope m of the linear

regressions to the hourly data, Ci = m · NOy and Fi = n · FNOy. Standard errors and p-

values for the regressions are given.

119

02

46

80

24

68

NW

0.0

0.4

0.8

1.2

0.0

0.4

0.8

1.2

0 5 10 15 20

-25

-15

-50

SW

0 5 10 15 20

Hour

Summer 2000

NOyNOx+HNO3+PANNOx+HNO3NOx = NO+NO2NO

Flux

(µm

ol m

-2hr

-1)

Con

cent

ratio

n (n

mol

mol

-1)

Frac

tion

of [N

Oy]

FPAN (est.)FNOx (param.)FHNO3 (DDIM)FNOy (e.c.)

Figure 4.5. Summer 2000 median concentrations (upper panels), fractions of NOy (middle

panels), and fluxes (lower panels) of NOy and component species separated by wind

direction (Northwest on the left, Southwest on the right). Vertical lines in the flux panels

show 25th and 75th quantiles of FNOy and FHNO3; negative fluxes represent deposition;

FPAN is an estimate, FNOx is parameterized, FHNO3 is inferred, and FNOy was

measured by eddy covariance. The NOy concentration and flux budgets are largely

closed for northwesterly (unpolluted background) flows, whereas up to 50% of NOy and

FNOy under southwesterly flows are in the form of reactive nitrogen species whose

fluxes are not measured or estimated here.

120

igure 4.6. Same format as Figure 4.5 for September, October, and November 2000.

PAN concentrations were not measured during the fall of 2000.

02

46

80

24

68

NW

0.0

0.4

0.8

1.2

0.0

0.4

0.8

1.2

0 5 10 15 20

-25

-15

-50

SW

0 5 10 15 20

Hour

Fall 2000

Flux

(µm

ol m

-2hr

-1)

Con

cent

ratio

n (n

mol

mol

-1)

Frac

tion

of [N

Oy]

NOyNOx+HNO3NOx = NO+NO2NO

FNOx (param.)FHNO3 (DDIM)FNOy (e.c.)

F

121

Table 4.1a,b. Fractional concentrations and fluxes of reactive nitrogen species at Harvard

Forest, 2000. Linear regressions (Ci=m·[NO]y and Fi=n·FNOy) were computed on hourly

data. PAN is not included in summations due to an insufficient number of overlapping

data hours. Day: 08:00-16:00 Eastern Standard Time. Night: 20:00-04:00 EST.

a) SUMMER NW SW

slope ± std. error (p-value)

day night day night

[NOx]/[NOy] 0.38±0.02 (<1E-4)

0.56±0.02 (<1E-4)

0.39±0.02 (<1E-4)

0.46±0.01 (<1E-4)

[HNO3]/[NOy] 0.35±0.05 (<1E-4)

0.28±0.04 (<1E-4)

0.20±0.02 (<1E-4)

0.19±0.02 (<1E-4)

[PAN]/[NOy] 0.133±0.008

(<1E-4) 0.131±0.005

(<1E-4) 0.138±0.009

(<1E-4) 0.069±0.005

(<1E-4)

([NOx]+[HNO3])/[NOy] 0.61±0.09 (3E-4)

0.8±0.2 (1E-3)

0.70±0.05 (<1E-4)

0.69±0.04 (<1E-4)

FNOx(dep)/FNOy 0.022±0.002

(<1E-4) 0.06±0.01 (<1E-4)

0.027±0.003 (<1E-4)

0.077±0.007 (<1E-4)

FHNO3/FNOy 0.7±0.1 (<1E-4)

0.2±0.2 (0.4)

0.46±0.06 (<1E-4)

0.6±0.1 (<1E-4)

FPAN/FNOy 0.039±0.003

(<1E-4) 0.096±0.006

(<1E-4) 0.047±0.005

(<1E-4) 0.049±0.005

(<1E-4)

(FNOx+FHNO3)/FNOy 0.8±0.1 (<1E-4)

0.2±0.1 (0.3)

0.53±0.04 (<1E-4)

0.5±0.1 (<1E-4)

b) FALL NW SW

slope ± std. error (p-value)

day night day night

[NOx]/[NOy] 0.83±0.05 (<1E-4)

0.65±0.04 (<1E-4)

0.73±0.02 (<1E-4)

0.58±0.03 (<1E-4)

[HNO3]/[NOy] 0.21±0.03 (<1E-4)

0.14±0.03 (<1E-4)

0.09±0.01 (<1E-4)

0.11±0.01 (<1E-4)

([NOx]+[HNO3])/[NOy] 1.03±0.07 (<1E-4)

0.97±0.06 (<1E-4)

0.84±0.02 (<1E-4)

0.70±0.03 (<1E-4)

FNOx(dep)/FNOy 0.08±0.01 (<1E-4)

0.21±0.03 (<1E-4)

0.13±0.03 (<1E-4)

0.18±0.03 (<1E-4)

FHNO3/FNOy 0.7±0.1 (<1E-4)

0.40±0.09 (1E-4)

0.74±0.07 (<1E-4)

0.38±0.07 (<1E-4)

(FNOx+FHNO3)/FNOy 0.9±0.2 (<1E-4)

0.6±0.1 (<1E-4)

0.9±0.1 (<1E-4)

0.6±0.1 (<1E-4)

122

During northwesterly flow conditions, the sum of individual reactive nitrogen

species, Σ[NOy]i = [NO] + [NO2] + [HNO3] + [PAN], was very close to the observed

concentration of NOy. HNO3 accounted for roughly a third of NOy during the day in

summer, a fifth in the fall, with the remainder of the NOy budget consisting largely of

NOx. The summertime concentration of PAN was on the order of 13% of [NOy]. During

northwesterly flows, NOy fluxes were dominated by HNO3 deposition fluxes in both

seasons.

During southwesterly flow conditions, a larger fraction of the NOy concentration

and flux was in the form of unmeasured species. HNO3 and NOx contributed less than

70% of the total NOy concentration, with summertime PAN accounting for 7 to 14%.

HNO3 flux was between 38 and 74% of FNOy, and neither FNOx nor FPAN accounted

for the difference.

4.4. Discussion

Accurate determination of the concentration of HNO3 depends critically on the

transmission efficiency through the inlet with varying temperature and humidity.

Because southwesterly winds at Harvard Forest tend to transport warmer, moister air than

northwesterlies [Moody et al., 1998], it is important to look closely at possible instrument

artifacts, especially of HNO3 transmission through the inlet, associated with these

differences. A complete description of TDLAS materials testing and response of the

fused silica and coated pyrex inlet at a typical ambient temperature and humidity is

included in Chapter 2. Neuman et al. [1999] noted a dramatic difference between the

transmission of HNO3 through Teflon fluoropolymer (“PFA”) tubing in dry and

123

humidified synthetic air. Ambient conditions at Harvard Forest rarely approach “dry”

laboratory conditions; during 90% of the summer 2000 measurement hours, relative

humidity was above 55%, and the minimum observed relative humidity during the

summer was 30.5%. In order to account for both temperature and relative humidity

differences between the dominant wind sectors, we used the H2O vapor pressure to

investigate a possible inlet response effect in the data. Mean ± standard deviation H2O

vapor pressure for the Northwest sector was 1.5 ± 0.4 KPa, and 1.9 ± 0.4 KPa for the

Southwest sector.

According to the HNO3 and NOy measurements presented in Section (4.3),

F(HNO3)/F(NOy) is smaller for southwesterly conditions where water vapor pressure is

higher. Inferred deposition velocities (from DDIM) for the two wind sectors are nearly

identical. In Figure 4.7, summer F(HNO3)/F(NOy) is shown as a function of H2O vapor

pressure for both wind sectors. Although the fractional HNO3 flux drops slightly at the

highest H2O vapor pressures observed under northwesterly conditions, it is consistently

lower under southwesterly conditions across all H2O vapor pressures. Note that the range

of observed vapor pressures overlaps considerably between the wind sectors. Although

we cannot rule out a small water vapor effect on HNO3 transmission in the TDLAS

instrument, the difference in F(HNO3)/F(NOy) between the wind sectors cannot be

entirely attributed to such an effect.

Catalytic NOy converters have been implicated in a number of cases where the

sum of individual reactive nitrogen species was substantially less than the observed

concentration of NOy. However, it is unlikely that this is the cause of the southwest

sector discrepancy observed at Harvard Forest. Kliner et al. [1997] found that conversion

124

01

23

40

12

34

0.0 0.5 1.0 1.5 2.0 2.5

01

23

4

0.0 0.5 1.0 1.5 2.0 2.5

01

23

4

FHN

O3/F

NO

yFH

NO

3/FN

Oy

H2O Vapor Pressure (KPa)

NW

SW

Figure 4.7. Fraction of HNO3 flux (inferred) to NOy flux (eddy covariance) as a function

of H2O vapor pressure for the northwest (upper) and southwest (lower) wind sectors at

Harvard Forest, summer 2000. Dots are individual hourly values, and squares with

vertical bars are means and standard deviations of the fractional flux binned by quantiles

of H2O vapor pressure.

125

of reduced nitrogen species HCN, CH3CN, and NH3 on a hot 24 k Au catalyst in the

presence of H2 (as used at Harvard Forest) was exacerbated after cleaning the catalyst and

under very low humidity conditions. The Harvard Forest catalyst is well-aged, and as

noted above, ambient conditions are never very dry. In addition, NH3 concentrations

during the summer at Harvard Forest are only on the order of several hundred parts per

trillion by volume ([NH3] ≈ 0.2 to 0.4 nmol mol-1), independent of wind direction [Lefer

et al., 1999], far too small to influence these results even if a portion of the NH3 were

oxidized to NO in the catalyst.

Particulate NO3- also contributes to the total NOy measurement, but was not

independently monitored during the summer of 2000. Lefer et al. [2001] found that

summer median values of aerosol nitrate between 1991 and 1995 were 4.92 nmol m-3 (≈

0.1 nmol mol-1) for the northwest wind sector, and 9.41 nmol m-3 (≈ 0.2 nmol

mol-1) for the southwest. The concentrations of particulate NO3- appear too low to

account for the observed southwest sector discrepancy.

Munger et al. [1996] defined the deposition velocity of (NOy-NOx), Vd(NOy-NOx)

= FNOy/([NOy]-[NOx]) and found that Vd(NOy-NOx) varied little between Northwest and

Southwest wind sectors at Harvard Forest, much like the inferred HNO3 deposition

velocities presented here. The species composing NOy-NOx were not measured in the

earlier work, so the fraction of rapidly-depositing species in NOy-NOx was not known.

Such efficient depositors could include HNO3 and organic hydrocarbon oxidation

products such as alkyl- and hydroxyalkyl nitrates. Munger et al. [1998] estimated the

production rates of hydroxyalkyl nitrates from biogenic emissions of isoprene and

126

monoterpenes at the site, and found that they could account for 25% of the observed NOy

deposition during the summer.

Alkyl nitrates, products of alkane oxidation, and hydroxyalkyl nitrates, products

of alkene oxidation, have been observed in both rural and urban air masses [Rosen et al.,

2001; Ostling et al., 2001; O’Brien et al., 1997; Shepson et al., 1993; O’Brien et al.,

1995]. The alkyl nitrates undergo dry deposition, and the hydroxyalkyl nitrates undergo

wet and dry deposition due to high Henry’s law coefficients [Shepson et al., 1996; Treves

et al., 2000]. The observed concentrations of these species at urban and rural sites have

been no larger than 3% of total NOy [Ostling et al., 2001; O’Brien et al., 1997; Shepson

et al., 1993; O’Brien et al., 1995], or 10-20% of (NOy-NOx) [Rosen et al., 2001].

As shown in Figure 4.8 for southwesterly flows during the summer of 2000, the

unmeasured fraction of [NOy], NOy-Σ(NOy)i where Σ(NOy)i = [NO] + [NO2] + [HNO3] +

[PAN], was between 1 and 3 nmol mol-1, or approximately 25% of [NOy]. The flux of

unmeasured NOy, FNOy-Σ(FNOy)i where Σ(FNOy)i = FNO + FNO2 + FHNO3 + FPAN, is

also shown in Figure 4.8. Deposition fluxes of NO and NO2 were derived from the

parameterization presented in Chapter 3, HNO3 flux was inferred as described above, and

the flux of PAN was estimated, also described above. The deposition velocity of

unmeasured NOy species, (FNOy-Σ(FNOy)i)/ (NOy-Σ(NOy)i) was approximately 3 cm s-1,

of the same order as HNO3 deposition velocities. Hydroxyalkyl nitrates are generally

assumed to deposit like HNO3, although we know of no independent determinations of

their deposition rates [e.g. Liang et al., 1998].

Overall, we find that HNO3 concentration and flux are not substantially different

between the clean and polluted wind sectors at Harvard Forest. The higher NOy

127

[NO

y-Σ(N

Oy) i

](nm

ol m

ol-1

)FN

Oy-Σ

(FN

Oy) i

(µm

ol m

-2hr

-1)

-Vd(

NO

y-Σ(

NO

y) i) (

cm s

-1)

Summer 2000, SW, NOy-Σ(NOy)i

01

23

45

01

23

45

-15

-10

-50

-15

-10

-50

0 5 10 15 20

-5-4

-3-2

-10

Hour

Figure 4.8. Estimated diel concentration (upper), flux (middle), and exchange velocity

(lower, negative sign represents deposition) of unmeasured NOy species for summer

2000, Harvard Forest. Σ(NOy)i = NO + NO2 + HNO3 + PAN. NO, NO2, HNO3, and

PAN concentrations were measured. NOx fluxes were parameterized based on measured

fluxes, HNO3 flux was inferred, and PAN flux was estimated. Mean ± standard deviation

unmeasured NOy deposition velocity for all hours was 2.9 ± 1.7 cm s-1.

128

deposition rates associated with southwesterly flow arise primarily from rapidly-

depositing, non-HNO3 species. Although the concentrations of our unmeasured species

exceed those of alkyl- and hydroxyalkyl nitrites observed in urban and rural locations, the

deposition velocity that we derive for these compounds is consistent with their assumed

deposition velocities. Their presence in the southwesterly flow regime suggests that

anthropogenic hydrocarbon emissions may be their source. Alternatively, the higher

temperatures and NOx concentrations during southwesterly flows may enhance the

formation of organic nitrates during the oxidation of biogenic hydrocarbons such as

isoprene and monoterpenes emitted from the forest.

It is not known whether the non-HNO3 depositors whose presence we have

inferred would produce a signal in filterpack, denuder, or mist chamber HNO3

measurements. Although dissolution to form NO3- in solution is possible, we do not

know if the efficiency for this process could be comparable to HNO3. These unidentified

species have potentially important ramifications for the total dry deposition of reactive

nitrogen that is presently estimated by filterpack HNO3 + particulate (NO3- + NH4

+)

measurements [e.g. Pratt et al., 1996]. Under conditions with large biogenic hydrocarbon

sources and moderate transport of anthropogenic NOx, the current estimates of dry

nitrogen deposition might be significantly in error.

4.5. Conclusions

We installed a Tunable Diode Laser Absorption Spectrometer (TDLAS) to

measure continuous HNO3 concentrations above the canopy at Harvard Forest during the

spring, summer, and fall of 2000. A conventional Dry Deposition Inferential Method

129

(DDIM) was employed to estimate the hourly deposition flux of HNO3. The diurnal

variation of HNO3 concentration is weak, and thus DDIM applied in a weekly average

sense to our hourly data, analogous to the practice at nitrogen deposition network sites,

introduced noise but not a persistent bias. Using new and ongoing NOy, NOx, and PAN

measurements at the site, we present speciated NOy concentration and flux budgets for

clean (northwesterly) and polluted (southwesterly) conditions during summer and fall.

HNO3 concentrations and fluxes differed little between the two flow regimes.

The concentration of PAN was low (<15% of NOy) and, with its small estimated

deposition velocity (0.8 ± 0.4 cm s-1), PAN contributed little to the flux of NOy.

Parameterized NOx deposition, based on eddy covariance NO and NO2 measurements at

the site [Chapter 3], showed that NOx played only a minor role in total NOy deposition.

For the clean (NW) sector, HNO3 accounted for nearly all of the NOy flux, and the

concentration budget for the northwesterly flow conditions was essentially closed using

data for NOx, HNO3, and PAN. For the more polluted, warmer, and moister southwest

sector, as much as 50% of the NOy concentrations and fluxes were unaccounted by the

measured species. The differences could not be explained by water vapor effects on the

HNO3 concentration measurement, NH3 interference in the NOy measurement, or

particulate NO3-. We believe that the deficit in NOy concentration corresponds to

unmeasured depositing species, and we infer associated deposition velocities similar to

HNO3, approximately 3 cm s-1. The unidentified compound(s) could play an important

role in removing reactive nitrogen from the troposphere, contribute substantially to

nitrogen deposition, and affect the nutrient and toxicity loading of vegetation locally

and/or regionally, depending on the exact nature of the substances. Identifying these

130

species and determining whether current nitrogen deposition monitoring techniques

detect them should be a priority.

References

Aneja, V. P., D.-S. Kim, M. Das, and B. E. Hartsell, Measurements and analysis of reactive nitrogen species in the rural troposphere of southeast United States: Southern Oxidant Study Site SONIA, Atmos. Environ 30, 649-659, 1996.

Bradshaw, J., S. Sandholm, R. Talbot, An update on reactive odd-nitrogen measurements made during recent NASA Global Tropospheric Experiment programs, J. Geophys. Res. 103, 19129-19148, 1998.

Bradshaw, J., D. Davis, G. Grodzinsky, S. Smyth, R. Newell, S. Sandholm, S. Liu, Observed distributions of nitrogen oxides in the remote free troposphere from the NASA Global Tropopsheric Experiment Programs, Reviews of Geophysics 38, 61-116, 2000.

Brook, J. R., F. Di-Giovanni, S. Cakmak, T. P. Meyers, Estimation of dry deposition velocity using inferential models and site-specific meteorology – uncertainty due to siting of meteorological towers, Atmos. Environ. 31, 3911-3919, 1997.

Clarke, J. F., E. S. Edgerton, B. E. Martin, Dry deposition calculations for the clean air status and trends network, Atmos. Environ. 31, 3667-3678, 1997.

Day, D. A., P. J. Wooldridge, M. B. Dillon, J. A. Thornton, R. C. Cohen, A thermal dissociation laser-induced fluorescence instrument for in situ detection of NO2, peroxy nitroates, alkyl nitrates, and HNO3, J. Geophys. Res. 107, 10.1029/2001JD000779, 2002.

DeMore, W. B., S. P. Sander, D. M. Golden, R. F. Hampson, M. J. Kurylo, C. J. Howard, A. R. Ravishankara, C. E. Kolb, M. J. Molina, Chemical kinetics and pohotochemical data for use in stratospheric modeling, evaluation 11, JPL publication 97-4, 1997.

Fehsenfeld, F. C., L. G. Huey, D. T. Sueper, R. B. Norton, E. J. Williams, Ground-based intercomparison of nitric acid measurement techniques, J. Geophys. Res. 103, 3343-3353, 1998.

Furutani, H. and H. Akimoto, Development and characterization of a fast measurement system for gas-phase nitric acid with a chemical ionization mass spectrometer in the marine boundary layer, J. Geophys. Res. 107, 10.1029/2000JD000269, 2002.

Harrison, R. M., J. L. Grenfell, S. Yamulki, K. C. Clemitshaw, S. A. Penkett, J. N. Cape, G. G. McFadyen, Budget of NOy species measured at a coastal site, Atmos. Environ. 33, 4255-4272, 1999.

131

Hanson, P. and S. E. Lindberg, Dry deposition of reactive nitrogen compounds: a review of leaf, canopy, and non-foliar measurements, Atmos. Environ. 25A, 1615-1634, 1991.

Hicks, B. B., D. D. Baldocchi, T. P. Meyers, R. P. Hosker, Jr., D. R. Matt, A preliminary multiple resistance routein for deriving dry deposition velocities from measured quantities, Water, Air, and Soil Pollution 36, 311-330, 1987.

Horii, C. V., M. Zahniser, D. Nelson, J. B. McManus, S. C. Wofsy, Nitric Acid and Nitrogen Dioxide Flux Measurements: a New Application of Tunable Diode Laser Absorption Spectroscopy, Proc SPIE 3758, 152-161, 1999.

Hosker, R. P. Jr., and S. E. Lindberg, Review: Atmospheric deposition and plant assimilation of gases and particles, Atmos. Environ. 16, 889-910, 1982.

Huey, L. G., E. J. Dunlea, E. R. Lovejoy, D. R. Hanson, R. B. Norton, F. C. Fehsenfeld, C. J. Howard, Fast time response measurements of HNO3 in air with a chemical ionization mass spectrometer, J. Geophys. Res. 103, 3355-3360, 1998.

Jacob, D. J., Heterogeneous chemistry and tropospheric ozone, Atmos. Environ. 34, 2131-2159, 2000.

Janson, R. and L. Granat, A foliar rinse study of the dry deposition of nitric acid to a coniferous forest, Agricultural and Forest Meteorology 98-99, 683-696, 1999.

Kliner, D. A. V., B. C. Daube, J. D. Burley, S. C. Wofsy, Laboratory investigation of the catalytic reduction technique for measurement of atmospheric NOy, J. Geophys. Res. 102, 10759-10776, 1997.

Lefer, B. L., The chemistry and dry deposition of atmospheric nitrogen at a rural site in the northeastern United States, Ph.D. thesis, 119 pp., Univ. of N. H., Durham, Dec. 1997.

Lefer, B. L., and R.W. Talbot, Summertime measurements of aerosol nitrate and ammonium at a northeastern U.S. site, J. Geophys. Res. 106, 20365-20378, 2001.

Lefer, B. L., and R. W. Talbot, Nitric acid and ammonia at a rural northeastern U.S. site, J. Geophys. Res. 104, 1645-1661, 1999.

Liang, J., L. W. Horowitz, D. J. Jacob, Y. Want, A. M. Fiore, J. A. Logan, G. M. Gardner, J. W. Munger, Seasonal budgets of reactivenitrogen species and ozone over the United States, and export fluxes to the global atmosphere, J. Geophys. Res. 103, 13425-13450, 1998.

Matt, D. R. and T. P. Meyers, On the use of the interential technique to estimate dry deposition of SO2, Atmos. Environ. 27A, 493-501, 1993.

132

Mauldin, R. L. III, D. J. Tanner, F. L. Eisele, A new chemical ionization mass spectrometer technique for the fast measurement of gas phase nitric acid in the atmosphere, J. Geophys. Res. 103, 3361-3367, 1998.

McFadyen, G. G., and J. Neil Cape, Physical and chemical influences on PAN concentrations at a rural site, Atmos. Environ. 33, 2929-2940, 1999.

Miller, C. E., J. I. Lynton, D. M. Keevil, J. S. Francisco, Dissociation Pathways of Peroxyacetyl Nitrate (PAN), J. Phys. Chem. A 103, 11451-11459, 1999.

Miller, T. M., J. O. Ballenthin, R. F. Meads, D. E. Hunton, W. F. Thorn, A. A. Viggiano, Y. Kondo, M. Koiki, Y. Zhao, Chemical ionization mass spectrometer technique for the measurement of HNO3 in air traffic corridors in the upper troposphere during the SONEX campaign, J. Geophys. Res 105, 3701-3707, 2000.

Meyers, T. P., B. J. Huebert, B. B. Hicks, HNO3 Deposition to a Deciduous Forest, Bounary-Layer Meteorology 49, 395-410, 1989.

Meyers, T. P., P. Kinkelstein, J. Clarke, T. G. Ellestad, P. F. Sims, A multilayer model for inferring dry deposition using standard meteorological measurements, J. Geophys. Res. 103, 22645-22661, 1998.

Moise, T. W. Denzer, Y. Rudich, Direct Kinetics Study of the Reaction of Peroxyacetyl Radical with No between 218 and 370 K, J. Phys. Chem. A. 103, 6766-6771, 1999.

Moody, J. L., J. W. Munger, A. H. Goldstein, D. J. Jacob, and S. C. Wofsy, Harvard Forest regional-scale air mass composition by Patterns in Atmospheric Transport History (PATH), J. Geophys. Res. 103, 13181-13194, 1998.

Munger, J. W., S. C Wofsy, P. S. Bakwin, S-M. Fan, M. L. Goulden, B. C. Daube, and A. H. Goldstein, Atmospheric deposition of reactive nitrogen oxides and ozone in a temperate deciduous forest and a subarctic woodland: 1. Measurements and mechanisms, J. Geophys. Res. 101, 12639-12657, 1996.

Munger, J. W., S-M. Fan, P. S. Bakwin, M. L. Goulden, A. H. Goltstein, A. S. Colman, and S. C. Wofsy, Regional budgets for nitrogen oxides from continental sources: Variations of rates for oxidation and deposition with season and distance from source regions, J. Geophys. Res. 103, 8355-8368, 1998.

Neuman, J. A., L. G. Huey, T. B. Ryerson, D. W. Fahey, Study of Inlet Materials for Sampling Atmospheric Nitric Acid, Environ. Sci. Technol. 33, 1133-1136, 1999.

O’Brien, J. M., P. B. Shepson, K. Murhuramu, C. Hao, H. Niki, D. R. Hastie, Measurements of alkyl and multifunctional organic nitrates at a rural site in Ontario, J. Geophys. Res. 100, 22795-22804, 1995.

O’Brien, J. M., P. B. Shepson, Q. Wu, T. Biesenthal, J. W. Bottenheim, H. A. Wiebe, K. G. Anlauf, P. Brickell, Production and distribution of organic nitrates, and their

133

relationship to carbonyl compounds in an urban environment, Atmos. Environ. 31, 2059-2069, 1997.

Ostling, K., B. Kelly, S. Bird, S. Bertman, M. Pippin, T. Thornberr, M. A. Carroll, Fast-turnaround alkyl nitrate measurements during the PROPHET 1998 summer intensive, J. Geophys. Res. 106, 24439-24449, 2001.

Parrish, D. D., M. P. Buhr,, M. Trainer, R. B. Norteon, J. P. Shimshock, F. C. Fehsenfeld et al., The total reactive oxidized nitrogen levels and the partitioning between the individual species at six rural sites in eastern North America, J. Geophys. Res. 98, 2927-2939, 1993.

Pratt, G. C., E. J. Orr, D. C. Bock, R. L. Strassman, D. W. Fundine, C. J. Twaroski, J. D. Thornton, T. P. Meyers, Estimation of Dry Deposition of Inorganics Using Filter Pack Data and Inferred Deposition Velocity, Environ. Sci. Techno. 30, 2168-2177, 1996.

Rosen, R. S., E. Wood, P. J. Wooldridge, J. A. Thornton, R. C. Cohen, E. J. Williams, F. C. Fehsenfeld, J. M. Roberts, S. R. Hall, R. E. Shetter, M. Martinez, H. Harder, W. H. Brune, The Chemistry of ΣANs, ΣPANs, HNO3, and O3 in the Houston Urban Plume, EOS Trans. AGU, 82, Fall Meeting Suppl., A51F-0118, 2001

Rosman, K., M. Shimmo, A. Karlsson, H.-C. Hansson, P. Keronen, A. Allen, G. Hoenninger, Laboratory and field investigations of a new and simple design for the parallel plate denuder, Atmos. Environ. 35, 5301-5310, 2001.

Schrimpf, W., K. Lienaerts, K. P. Müller, J. Rudolph, R. Neubert, W. Schüßler, I. Levin, Dry deposition of peroxyacetyl nitrate (PAN): Determination of its deposition velocity at night from measurements of the atmospheric PAN and 222Radon concentration gradient, Geophys. Res. Lett. 23, 3599-3602, 1996.

Shepson, P. B., J. W. Bottenheim, D. R. Hastie, A. Venkatram, Determination of the relative ozone and PAN deposition velocities at night, Geophys. Res. Lett. 19, 1121-1124, 1992.

Shepson, P. B., K. G. Anlauf, J. W. Bottenheim, H. A. Wiebe, N. Gao, K. Muthuramu, G. I. Mackay, Alkyl nitrates and their contribution to reactive nitrogen at a rural site in Ontario, Atmos. Environ. 27A, 749-757, 1993.

Shepson, P. B., E. Mackay, K. Muthuramu, Henry’s Law Constants and Removal Processes for Several Atmospheric β-Hydroxy Alkyl Nitrates, Environ. Sci. Technol. 30, 3618-3623, 1996.

Sievering, H. T. Kelly, G. McConville, C. Seibold, A. Turnipseed, Nitric acid dry deposition to conifer forests: Niwot Ridge spruce-fir-pine study, Atmos. Environ. 35, 3851-3859, 2001.

134

135

Talbot, R. W., J. E. Dibb, B. L. Lefer, E. M. Scheuer, J. D. Bradshaw, S. T. Sandholm, S. Smyth, D. R. Blake, N. J. Blake, G. W. Sachse, J. E. Collins, G. L. Gregory, Large-scale distributions of tropospheric nitric, formic, and acetic acids over the western pacific basin during wintertime, J. Geophys. Res. 102, 28303-28313, 1997.

Thornberry, T. M. A. Carroll, G. J. Keeler, S. Sillman, S. B. Bertman, M. R. Pippin K. Ostling, J. W. Grossenbacher, P. B. Shepson, O. R. Cooper, J. L. Moody, W. R. Stockwell, Observations of reactive oxidized nitrogen and speciation of NOy during the PROPHET summer 1998 intensive, J. Geophys. Res. 106, 24359-24386, 2001.

Treves, K., L. Shragina, Y. Rudich, Henry’s Law Constants of some β-, γ-, and δ-Hydroxy Alkyl Nitrates of Atmospheric Interest, Environ. Sci. Technol. 34, 1197-1203, 2000.

U.S. Environmental Protection Agency, Atmospheric Deposition of Sulfur and Nitrogen Compounds, in National Air Quality and Emission Trends Report, 1999, EPS454/R-01-004, Office of Air Quality Planning and Standards, Research Triangle Park, NC, 2001.

Warneck, P. and T. Zerbach, Synthesis of peroxyacetyl nitrate in air by acetone photolysis, Environ. Sci. Technol. 26, 74-79, 1992.

Wesely, M. L. and B. B. Hicks, Some Factors that Affect the Deposition Rates of Sulfur Dioxide and Similar Gases on Vegetation, Journal of the Air Pollution Control Association 27, 1110-1116, 1977.

Wesely, M. L., Parameterization of surface resistances to gaseous dry deposition in regional-scale numerical models, Atmos. Environ. 23, 1293-1304, 1989.

Wesley, M. L., and B. B. Hicks, A review of the current status of knowledge on dry deposition, Atmos. Environ. 34, 2261-2282, 2000.