ArticlePDF Available

The T1-T2 study: Evolution of aerosol properties downwind of Mexico City

Authors:
  • Olallie Laboratory Inc.

Abstract and Figures

As part of a major atmospheric chemistry and aerosol field program carried out in March 2006, a study was conducted in the area to the north and northeast of Mexico City to investigate the evolution of aerosols and their associated optical properties in the first few hours after their emission. The focus of the T1-T2 aerosol study was to investigate changes in the specific absorption α<sub>ABS</sub> (absorption per unit mass, with unit of m<sup>2</sup> g<sup>−1</sup>) of black carbon as it aged and became coated with compounds such as sulfate and organic carbon, evolving from an external to an internal mixture. Such evolution has been reported in previous studies. The T1 site was located just to the north of the Mexico City metropolitan area; the T2 site was situated approximately 35 km farther to the northeast. Nephelometers, particle soot absorption photometers, photoacoustic absorption spectrometers, and organic and elemental carbon analyzers were used to measure the optical properties of the aerosols and the carbon concentrations at each of the sites. Radar wind profilers and radiosonde systems helped to characterize the meteorology and to identify periods when transport from Mexico City over T1 and T2 occurred. Organic and elemental carbon concentrations at T1 showed diurnal cycles reflecting the nocturnal and early morning buildup from nearby sources, while concentrations at T2 appeared to be more affected by transport from Mexico City. Specific absorption during transport periods was lower than during other times, consistent with the likelihood of fresher emissions being found when the winds blew from Mexico City over T1 and T2. The specific absorption at T2 was larger than at T1, which is also consistent with the expectation of more aged particles with encapsulated black carbon being found at the more distant location. In situ measurements of single scattering albedo with an aircraft and a ground station showed general agreement with column-averaged values derived from rotating shadowband radiometer data, although some differences were found that may be related to boundary-layer evolution.
Content may be subject to copyright.
Atmos. Chem. Phys., 7, 1585–1598, 2007
www.atmos-chem-phys.net/7/1585/2007/
© Author(s) 2007. This work is licensed
under a Creative Commons License.
Atmospheric
Chemistry
and Physics
The T1-T2 study: evolution of aerosol properties downwind of
Mexico City
J. C. Doran1, J. C. Barnard1, W. P. Arnott2, R. Cary3, R. Coulter4, J. D. Fast1, E. I. Kassianov1, L. Kleinman5,
N. S. Laulainen1, T. Martin4, G. Paredes-Miranda2, M. S. Pekour1, W. J. Shaw1, D. F. Smith3, S. R. Springston5, and
X.-Y. Yu1
1Pacific Northwest National Laboratory, Richland, WA, USA
2Desert Research Institute, Reno, NV, USA
3Sunset Laboratory, Inc., Tigard, OR, USA
4Argonne National Laboratory, Argonne, IL, USA
5Brookhaven National Laboratory, Upton, NY, USA
Received: 6 November 2006 – Published in Atmos. Chem. Phys. Discuss.: 12 December 2006
Revised: 22 February 2007 – Accepted: 15 March 2007 – Published: 23 March 2007
Abstract. As part of a major atmospheric chemistry and
aerosol field program carried out in March 2006, a study was
conducted in the area to the north and northeast of Mexico
City to investigate the evolution of aerosols and their associ-
ated optical properties in the first few hours after their emis-
sion. The focus of the T1-T2 aerosol study was to investi-
gate changes in the specific absorption αABS (absorption per
unit mass, with unit of m2g1) of black carbon as it aged
and became coated with compounds such as sulfate and or-
ganic carbon, evolving from an external to an internal mix-
ture. Such evolution has been reported in previous studies.
The T1 site was located just to the north of the Mexico City
metropolitan area; the T2 site was situated approximately
35km farther to the northeast. Nephelometers, particle soot
absorption photometers, photoacoustic absorption spectrom-
eters, and organic and elemental carbon analyzers were used
to measure the optical properties of the aerosols and the car-
bon concentrations at each of the sites. Radar wind profilers
and radiosonde systems helped to characterize the meteorol-
ogy and to identify periods when transport from Mexico City
over T1 and T2 occurred. Organic and elemental carbon con-
centrations at T1 showed diurnal cycles reflecting the noc-
turnal and early morning buildup from nearby sources, while
concentrations at T2 appeared to be more affected by trans-
port from Mexico City. Specific absorption during transport
periods was lower than during other times, consistent with
the likelihood of fresher emissions being found when the
winds blew from Mexico City over T1 and T2. The spe-
Correspondence to: J. C. Doran
(christopher.doran@pnl.gov)
cific absorption at T2 was larger than at T1, which is also
consistent with the expectation of more aged particles with
encapsulated black carbon being found at the more distant lo-
cation. In situ measurements of single scattering albedo with
an aircraft and a ground station showed general agreement
with column-averaged values derived from rotating shadow-
band radiometer data, although some differences were found
that may be related to boundary-layer evolution.
1 Introduction
The MILAGRO (Megacity Initiative: Local and Global
Research Observations; http://www.eol.ucar.edu/projects/
milagro/) field campaign was designed to follow the urban
plume originating in Mexico City in order to study the evo-
lution of the properties of trace gases and aerosols as they
drifted downwind from a megacity. The study was con-
ducted over multiple scales, ranging from ground-based in-
vestigations centered in the Mexico City metropolitan area
to aircraft sampling over distances of hundreds of kilome-
ters. A major component of the MILAGRO campaign was
the MAX-MEX experiment (Megacity Aerosol Experiment
in Mexico City; http://www.asp.bnl.gov/MAX-Mex.html),
which focused on measurements within, over, and a few tens
of kilometers downwind from Mexico City. MAX-MEX was
supported by the U.S. Department of Energy through its At-
mospheric Sciences Program. In this paper we describe some
aspects of one element of the MAX-MEX experiment that we
refer to as the T1-T2 study.
Published by Copernicus GmbH on behalf of the European Geosciences Union.
1586 J. C. Doran et al.: The T1-T2 study
The T1-T2 study was motivated by the desire to investi-
gate the evolution of aerosols and their associated optical
properties in the first few hours after their formation. We
hypothesized that the shift from an externally mixed state
characteristic of fresh emissions to an internally mixed one
would result in significant modifications to the optical prop-
erties of aerosols as they were advected downstream from
Mexico City. Evidence for changes in aerosol composition
due to aging in the Mexico City area has been found by other
investigators (e.g., Baumgardner et al., 2000; Johnson et al.,
2005). We were particularly interested in changes in the mass
absorption efficiency or specific absorption αABS (absorption
per unit mass, with unit of m2g1)of black carbon (BC) be-
cause of its potential importance for climate change. The
specific absorption is one of two key parameters that deter-
mine the magnitude of the aerosol forcing by BC, the other
being the atmospheric burden (Chung and Seinfeld, 2002;
Sato et al., 2003). In this paper we will use the terms black
carbon and elemental carbon (EC) interchangeably, but we
note here that Andreae and Gelencs´
er (2006) warn that the
term “black carbon” has been used somewhat carelessly in
the literature. For the moment we will not deal with this dif-
ficulty but will return to it when we discuss our absorption
measurements in Sect. 5.
In the past, field estimates of the specific absorption of
black carbon have varied by an order of magnitude, ranging
from 2m2g1to 25m2g1at 500nm. (e.g., Waggoner
et al., 1981; Horvath, 1993; Liousse et al., 1993; Petzold et
al., 1997; Penner et al., 1998; Moosm¨
uller, 1998; Marley et
al., 2001; Baumgartner, 2002; Arnott et al., 2003; Schnaiter
et al., 2003; Schuster et al., 2005; Mikhailov et al., 2006).
While BC is normally assumed to be present as an external
mixture close to the source (e.g., Mallet et al., 2003; Jacob-
son and Seinfeld, 2004), processes such as heterocoagulation
and condensation can alter the mixing state as the particles
age (e.g., Jacobson, 2001). Such processes likely have con-
tributed to the order of magnitude differences in the estimates
of αABS noted above. Recently, Bond and Bergstrom (2006)
suggested that freshly emitted, uncoated soot particles have a
value of αABS=7.5±1.2 m2g1at 550 nm. This “uncoated”
value will increase as the particles age and become coated,
e.g., with substances such as sulfate or organic carbon, but
the rates at which aging occurs and direct observations of
the change in αABS as this occurs downstream from a source
have been difficult to determine in the field. In contrast, the
theoretical study of Fuller et al. (1999) suggests that αABS is
unlikely to exceed 10 m2g1unless mostof the BC is encap-
sulated in aerosol distributions whose geometric mean radius
is greater than 0.06µm. In light of these and other studies,
we are led to ask: to what extent and at what rate does aging
affect αABS for BC-containing aerosols that were originally
uncoated?
The campaign in the Mexico City area provided an op-
portunity to address this question by measuring values of
αABS at two or more locations, one relatively close to ma-
jor sources of BC and one farther downstream. Unless an
aerosol is somehow tagged to a specific source, the precise
location and time of the formation, t0, of an aerosol over
Mexico City is essentially impossible to determine after it
has drifted downwind by a few kilometers. Once outside the
primary source region, however, it is possible, at least in prin-
ciple, to track air masses containing aerosols. Our approach
was to measure aerosol properties at two sites downwind of
the Mexico City urban area and to use meteorological mea-
surements and modeling to determine periods when aerosols
originating in Mexico City passed over the two sites. During
those periods a segment of the urban plume can be assumed
to arrive at the first downwind location at a time t1and at the
second location at a time t2, thereby giving rise to the T1-T2
identifying label for this study. The differences in aerosol
properties as a function of the time difference t2–t1can then
be studied.
In the remainder of this paper we describe the design of the
T1-T2 study, compare some of the meteorological conditions
found over the T1 and T2 sites, identify periods when the
Mexico City plume was expected to flow over both sites, and
present some preliminary comparisons of specific absorption
and other aerosol properties at the two locations. More exten-
sive and detailed comparisons, as well as analyses of results
obtained at T0, will be the subjects of future papers.
2 Study design
The T1 site was located at Tecamac University at 19.703N
latitude and 98.982W longitude, at an altitude of 2273 m.
This site is near built-up regions and a major highway
and is just to the north of the main metropolitan area
of Mexico City. Because the site is close to the city
we expected it would frequently be affected by the ur-
ban plume drifting over the region. An analysis com-
bining computer modeling, radar wind profiler and ra-
diosonde data collected during the 1997 IMADA-AVER ex-
periment (Doran et al., 1998), and rawinsonde data from
2002 and 2003 (http://mirage-mex.acd.ucar.edu/Documents/
MIRAGE-Mex SOD 040324.pdf) indicated that, in the up-
per portions of the boundary layer and above, flows from
Mexico City to the north, northeast, or northwest could be ex-
pected roughly 50% of the time during the month of March.
Closer to the surface the winds tend to be lighter and less
well organized. It was also estimated that winds throughout
the depth of the boundary layer would carry the Mexico City
urban plume to the north-northeast over the city of Pachuca
between 20 and 30% of the time. Additional information
about the meteorology encountered during the MILAGRO
campaign will be published elsewhere.
Although the expected frequency of favorable wind direc-
tions was not as high as we might have wished, it was de-
cided nonetheless that it would still be worthwhile to locate
the T2 site in the general area of Pachuca. Accordingly, the
Atmos. Chem. Phys., 7, 1585–1598, 2007 www.atmos-chem-phys.net/7/1585/2007/
J. C. Doran et al.: The T1-T2 study 1587
Fig. 1. Map of study area showing locations of T1 and T2 sites.
Topography contour intervals are 200 m.
T2 site was situated 35 km to the north-northeast of T1 at
Rancho la Bisnaga, 20.010 N latitude and 98.909 W longi-
tude, at an altitude of 2542 m. The locations of these two
sites are shown in Fig. 1. The region between T1 and T2
is much less densely developed than the Mexico City urban
area. Some local sources of BC and other aerosols can be
found between T1 and T2, primarily along the highway be-
tween Mexico City and Pachuca, but their impact was ex-
pected to be small compared to that of the urban plume orig-
inating in Mexico City. Also shown in Fig. 1 is the location of
the Insituto Mexicano del Petr´
oleo, labeled T0 in the figure,
which was an observation site considered to be representative
of the Mexico City metropolitan area.
The T1 site was one of the two major ground sites for
the MILAGRO campaign, with investigators from approxi-
mately a dozen laboratories and universities making a wide
range of meteorological, chemical, aerosol, and atmospheric
radiation measurements. In contrast, the T2 site featured a
much smaller number of investigators and instruments, con-
sistent with the expected infrequency of urban plume pas-
sages over that location. In this paper we will consider data
from pairs of instruments, with one of each pair located at
both T1 and T2: radar wind profilers, radiosondes, organic
and elemental carbon (OCEC) analyzers, photoacoustic ab-
sorption spectrometers (PASs), 3-wavelength nephelometers,
particle soot absorption photometers (PSAPs), and multi-
filter rotating shadowband radiometers (MFRSRs). Informa-
tion about these instruments is provided below. In addition to
these ground-based instruments, several aircraft made flights
over T1 and T2 during the campaign. We will briefly dis-
Fig. 2. Time series of temperature, humidity, and total solar radia-
tion measured with surface instruments located at T2.
cuss some results from measurements with a nephelometer
and PSAP on board DOE’s Gulfstream-1 (G-1) aircraft taken
during periods when the Mexico City plume was expected to
drift over those sites.
Although some instruments were operational earlier, full-
scale operations at T1 and T2 commenced approximately on
day of year (DOY) 70 (11 March). Measurements at both
sites continued from that time through DOY 87 (28 March).
Figure 2 shows time series of temperature, humidity, and to-
tal solar radiation measured with surface instruments located
at T2. The first two weeks of the measurement period were
generally warmer, drier, and had less cloudiness than the last
week. Rain showers occurred on a daily basis during the last
week.
3 Instruments
The radar wind profilers (RWPs) used at T1 and T2 were
915MHz models manufactured by Vaisala. They were oper-
ated in a 5-beam mode with nominal 192-m range gates. We
applied the NCAR Improved Moment Algorithm (Morse et
al., 2002) to the moments data from the profilers to obtain
30-min average consensus winds. The radiosonde systems
were Vaisala DigiCORAsystems and used model RS-92
www.atmos-chem-phys.net/7/1585/2007/ Atmos. Chem. Phys., 7, 1585–1598, 2007
1588 J. C. Doran et al.: The T1-T2 study
sondes, also manufactured by Vaisala. The sondes released
at T2 measured temperature and humidity while some of the
sondes released at T1 also measured winds with a GPS sys-
tem. On days when the G-1 was expected to fly in the vicinity
of T1, up to five radiosondes were released there, nominally
at 09:00, 11:00, 13:00, 15:00, and 17:00 local standard time
(LST). If the anticipated G-1 flights included sampling in the
T2 area, up to three additional sondes were launched at T2 at
11:00, 13:00, and 15:00 LST.
The OCEC analyzers were manufactured by Sunset Labo-
ratory Inc., and are similar to the thermal-optical instrument
described in Birch and Cary (1996). Samples collected on
a quartz filter are volatilized under controlled conditions of
temperature and processing gas mixtures. The process is
used to differentiate between organic and elemental carbon,
with an optical correction technique applied to correct for
pyrolytically generated carbon. Sampling times were 45 min
in length, with an additional 15 min allocated to analysisand
cooling after each sampling period. The OCEC unit was cal-
ibrated on-site using external standard filters before ambi-
ent sampling. The external standard was analyzed in off-line
mode. The data acquisition parameters were adjusted after
the calibration to reflect the known OC/EC ratio. The esti-
mated detection limit is 0.02 µgC m3and the estimated un-
certainty of the OCEC measurement is 0.2 µgC m3. Quartz
filters were changed every few days before a significant re-
duction in the intensity of the laser signal used for the optical
corrections was observed. A blank sample was scheduled
for midnight (LST) daily. We restricted our analyses to pe-
riods when the OC and EC concentrations were greater than
or equal to 0.1µgC m3.
The PAS measures sound pressure produced by light ab-
sorption in an acoustic resonator and has been described by
Arnott et al. (1999). The instruments at T1 and T2 operated
at a wavelength of 870nm. Operation at 870 nm is advan-
tageous because BC does absorb at this wavelength but the
absorption of other atmospheric constituents such as dust and
OC is quite weak (Jacobson, 1999; Sokolik and Toon, 1999;
Sato et al., 2003; Kirchstetter et al., 2004). Average absorp-
tion values were computed every two minutes, and we com-
bined absorption values averaged over 46-min periods with
EC values from the OCEC analyzers to obtain estimates of
the specific absorption of EC at T1 and T2.
The nephelometers (TSI 3563) and PSAPs (Radiance Re-
search) at T1 and T2 were configured as parts of aerosol op-
tical measurement systems developed by NOAA’s Climate
Modeling & Diagnostics Laboratory (CMDL). Ambient air
is diverted alternately through 1µm and 10µm impactors at
6-min intervals, and the humidity is controlled so that it is
less than or equal to 40% RH before the air is introduced
into the nephleometer or PSAP. The nephelometers oper-
ated at three wavelengths: 450 nm, 550nm, and 700nm; the
PSAPs also operated at three wavelengths: 470 nm, 530nm,
and 660nm. Scattering and absorption measurements were
averaged and recorded every minute. The nephelometer
and PSAP on the G-1 operated at the same wavelengths as
the surface units. Corrections to the PSAP absorption and
the nephelometer scattering were carried out using the ap-
proach described by Bond et al. (1999) and Anderson and
Ogren (1998). Additional information on the nephelome-
ters and PSAPs can be found at the CMDL website http:
//www.cmdl.noaa.gov/aero/instrumentation/instrum.html.
The MFRSRs (Harrison et al., 1994) measured the direct
normal, diffuse horizontal, and total horizontal components
of the broadband irradiances using a silicon detector, and
these same components at six wavelengths (415, 500, 615,
673, 870, and 940 nm) with an interference filter/silicon de-
tector combination having a nominal passband of 10nm at
each wavelength. Data were averaged and recorded at 20-
s intervals. From these measurements, we can determine
aerosol optical thickness, τ. By applying the technique of
Kassianov et al. (2005) it is also possible to determine the
aerosol single scattering albedo, $0, and asymmetry param-
eter, g. In contrast to the aerosol quantities measured by our
other surface instruments, these quantities should be consid-
ered as averages over the atmospheric column.
4 Boundary-layer structure and winds
Doran et al. (1998) found that on clear days the bound-
ary layer (BL) in Mexico City during March typically grew
slowly after sunrise (approximately 06:40 LST) to a depth
on the order of 1000m by 11:00 or 12:00 LST, after which
it grew rapidly to heights of 3000m or more in the next few
hours. This behavior was observed at T1 during the MILA-
GRO campaign as well, although the heights tended to be
somewhat smaller during this campaign, and the BL behav-
ior at the T1 and T2 sites appeared to be similar. Figure 3
shows a comparison of BL depths estimated from potential
temperature profiles measured with radiosondes at 11:00,
13:00, and 15:00 LST on days for which nearly simulta-
neous releases were made at the two sites. The regression
line is given by h2=0.90×h1+25m, with R2=0.89, where h1
and h2 are the BL depths at T1 and T2, respectively. We
have also compared boundary-layer depths at T1 and T2 by
examining the signal-to-noise ratio of the reflected signals
from the RWPs at the two sites (not shown). The agreement
between the two sites is again reasonable (R2=0.90), with
h2=0.79×h1+260m. The similarity of the boundary layer
depths at the two sites simplifies some aspects of the anal-
ysis. It makes plume trajectory analyses easier (see below)
and it helps ensure that aircraft operations in the mixed layer
at one site will also be in the mixed layer at the second site,
providing the flight altitudes are not too different. Further in-
formation on boundary layer characteristics during the cam-
paign will be the subject of other papers.
We were particularly interested in days when the winds
were likely to blow the urban plume over both T1 and T2.
A detailed analysis identifying times when this occurred will
Atmos. Chem. Phys., 7, 1585–1598, 2007 www.atmos-chem-phys.net/7/1585/2007/
J. C. Doran et al.: The T1-T2 study 1589
be carried out in the future using a mesoscale model that in-
corporates a data assimilation scheme, parameterizations of
vertical mixing, and estimates of source locations in the Mex-
ico City metropolitan area. For this initial analysis, however,
we have simply used the wind data obtained from the RWP
at T1 to calculate forward and back trajectories of air masses
that passed over T1. Figure 4 shows examples of trajectories
calculated at 1000m above ground level (AGL) for daylight
hours over a 20-day period. The most favorable conditions
for T1-T2 transport are seen to have occurred on DOY 69,
77, 78, 79, and 83, (10, 18, 19, 20, and 24 March, respec-
tively) with briefer periods of preferred wind directions oc-
curring on several other days. On some days (e.g., DOY 81
[22 March]) the calculated trajectories showed the Mexico
City plume traveled close to T2, and lateral spreading of the
plume may well have resulted in some impact of Mexico City
emissions on both T1 and T2. On other days (e.g., DOY 74
[14 March]), the trajectories show near-misses at T2 but the
back trajectories at T1 suggest that the air mass traveling
from T1 to T2 did not originate over the urban region. In
some cases the wind directions exhibited substantial vertical
shear (not shown), which makes interpretation of the plume
transport problematic.
On many days, and especially on those with wind direc-
tions favorable for T1-T2 transport, the boundary layer was
quite dry. For example, by 11:00 LST on each of DOY 70,
77, 78, and 79 (10, 18, 19, and 20 March) radiosonde sound-
ings at T1 showed relative humidities in the boundary layer
of 46% or less (usually much less), with similar values found
from the T2 sondes. Above the boundary layer there was
often a very dry layer that could extend over depths of hun-
dreds of meters or more. Figure 5 shows an example of this
feature. Hygroscopic growth of aerosols and the consequent
effects on aerosol light scattering can be assumed to be es-
sentially negligible under such dry conditions (Nemesure et
al., 1995; Baumgardner and Clarke, 1998; Im et al., 2001;
Redemann et al., 2001; Carrico et al., 2003; Markowicz et
al., 2003). Thus, the optical properties of aerosols measured
near the surface are likely to be similar to those aloft in a
well-mixed boundary layer, which is advantageous for some
radiative closure analyses.
5 Results and discussion
5.1 OC and EC concentrations
Figure 6 shows plots of the organic (OC) and elemental (EC)
carbon concentrations at T1 and T2 as a function of time. At
T1 there was often a strong diurnal variation in OC and EC
concentrations during the first two weeks, with concentra-
tions increasing during the nighttime hours and the largest
values occurring in the morning hours around sunrise or
shortly after. This behavior is consistent with the trapping
of pollutants in the Mexico City area overnight and during
Fig. 3. Comparison of boundary-layer depths at T1 and T2 esti-
mated from potential temperature profiles measured with radioson-
des at 11:00, 13:00, and 15:00 LST on days for which nearly
simultaneous releases were made at the two sites: 11:00 LST
(DOY 68, 74, 77-78), 13:00 LST (DOY 68, 69, 76–79), and
15:00 LST (DOY 68 and 79).
the morning hours in a shallow surface layer before the rapid
growth of the mixed layer commences in the late morning.
Although the boundary layer structure and evolution at T2
were similar to those at T1, the OC and EC behavior at T2
was much less regular; there were often multiple peaks dur-
ing a 24-h period, and mid- to late-afternoon peaks were
common. No significant accumulations of pollutants, com-
parable to what was found at T1, developed overnight at T2.
This was expected, given the absence of local sources in the
area around and immediately upwind of T2.
Transport from Mexico City to T2 appears to account for a
substantial fraction of the variations in OC and EC concentra-
tions at T2. OC and EC concentrations were generally lowest
at both sites during the last week of the campaign when the
weather conditions were most unsettled and there were peri-
ods of rain. On days with the most favorable wind directions
for T1-T2 transport (DOY 69, 77, 78, and 79 (10, 18, 19, and
20 March)), both the OC and EC values at T2 tended to reach
higher values than on other occasions, although DOY 70, 71,
and 81 (11, 12 and 22 March) also have elevated readings. At
T1, which was considerably closer to the local and Mexico
City sources than T2, the carbon concentrations showed less
sensitivity to wind directions.
DOY 83 (24 March) has been identified as a transport day
but shows some unusual features. The OC and EC concen-
trations at T1 were low throughout the night, but the OC val-
ues were considerably higher during the day while the EC
values were not. At T2, OC and EC were low at night but
both increased substantially shortly after sunrise. Rain fell
at T1 and T2 just after sunset the previous night and during
www.atmos-chem-phys.net/7/1585/2007/ Atmos. Chem. Phys., 7, 1585–1598, 2007
1590 J. C. Doran et al.: The T1-T2 study
Fig. 4. Trajectories of air parcels passing over T1 at 1000m a.g.l. for the hours 06:00–18:00 LST based on 30-min RWP data at T1. Blue
indicates flow into T1 and red indicates flow away from T1. The small box around T2 is 10km on a side; the larger box around T0 is an
approximate boundary of Mexico City.
Atmos. Chem. Phys., 7, 1585–1598, 2007 www.atmos-chem-phys.net/7/1585/2007/
J. C. Doran et al.: The T1-T2 study 1591
Fig. 5. Example of relative humidity profiles at T1 and T2 showing dry layer above the boundary layer for sounding taken at 13:00 LST
on DOY 78 (19 March). The stair-step pattern at T1 is due to the limited resolution of the humidity measurements for the system deployed
there.
the afternoon of DOY 83. The rain may have contributed to
the lower carbon concentrations and the specific absorption
values are somewhat higher than the median values found
for other transport periods. While this is consistent with the
results found by Redemann et al. (2001) and Mikhailov et
al. (2006), the sampling period is rather small and we hesi-
tate to attach too much significance to this result at this time.
There is considerable scatter about the means, but for the
period shown in Fig. 6 the OC concentrations at T2 were
approximately 23% lower than at T1 while the EC concen-
trations were smaller by almost a factor of four. The average
EC/OC ratio at T1 was 0.210, while that at T2 was approxi-
mately three times smaller with a value of 0.063. Assuming
the Mexico City metropolitan area is the primary source of
EC in the area, the decrease in this quantity between T1 and
T2 was anticipated, given the dilution expected in traveling
between the two sites. Conversely, the much smaller change
in OC concentrations presumably reflects the rapid and ongo-
ing generation of OC over the region (Volkamer et al., 2006).
5.2 Specific absorption of EC
There is still some debate about exactly what is measured by
an OCEC analyzer. Andreae and Gelencs´
er (2006) propose
an operational definition for apparent elemental carbon, as
“the fraction of carbon that is oxidized above a certain tem-
perature threshold in the presence of an oxygen-containing
atmosphere”, and denote this quantity as ECa. For simplic-
ity in this paper we have retained the more commonly used
designation EC, but we are mindful of this potentially im-
portant distinction. Under some circumstances EC may only
be loosely related to soot particles, but the connection should
be considerably better in regions with high concentrations of
petroleum combustion sources, and Mexico City is one such
region. Andreae and Gelencs´
er also note that BC is often
used to denote the result of a light-absorbing carbon mea-
surement by an optical absorption technique. In any event,
our interest for this study was not the measurement of the
specific absorption of soot per se but rather the absorption
www.atmos-chem-phys.net/7/1585/2007/ Atmos. Chem. Phys., 7, 1585–1598, 2007
1592 J. C. Doran et al.: The T1-T2 study
Fig. 6. Time series of the organic (OC) and elemental (EC) carbon concentrations at T1 and T2. Nighttime periods are indicated by the
shaded areas. Note differences in scales for T1 and T2.
per unit mass of EC and how that may be modified as EC be-
comes coated in the hours or days after emission. We again
note that almost all of the absorption measured by the PASs
at T1 and T2 is likely due to the EC at the respective sites
because dust and OC do not absorb significantly at the wave-
length of the PAS measurements. Figure 7 shows time series
of EC concentrations, absorption measured by the PASs, and
specific absorption computed from the ratio of the absorp-
tion determined from the PASs and the EC concentrations
measured by the OCEC analyzers, for a 12-day period dur-
ing which the OCEC analyzers and PASs at both sites had
good data recovery.
Figure 8 shows a scatter plot of the absorption as a func-
tion of EC concentrations at the two sites. There is a greater
degree of scatter at T2 than at T1, which is expected given the
substantially lower EC concentrations and absorption found
at the former site. At T1 the EC concentrations and absorp-
tion values were as much as an order of magnitude higher
than at T2 but the specific absorption was generally higher at
T2 than at T1.
The figure shows that reasonably robust fits to the data at
each site can be obtained with linear fits that are constrained
to go through the origin. The R2values in each case are
0.96. The slopes of the lines in the figure correspond to
the mean values of specific absorption at the two sites, with
the value at T2 about 9% higher than that at T1 for the full
sampling period. One can also obtain a least squares fit to
the data at each site while leaving the intercept as a free
Atmos. Chem. Phys., 7, 1585–1598, 2007 www.atmos-chem-phys.net/7/1585/2007/
J. C. Doran et al.: The T1-T2 study 1593
Fig. 7. Time series of EC concentrations, absorption measured by
the PASs, and specific absorption of EC at T1 and T2 for a 12-day
period during which the OCEC analyzers and PASs at both sites had
good data recovery. Transport periods are indicated by the shaded
areas.
parameter. The results are virtually unchanged at T1 but
result in a smaller slope at T2 and an intercept on the or-
der of 0.6Mm1. A non-zero intercept is unphysical (if the
EC concentration is zero the absorption should be zero) and
may indicate errors in the measurements, another absorbing
species, or that the specific absorption is not necessarily in-
dependent of concentration. At this time we have no way of
distinguishing among these possibilities. Partly for this rea-
son, and partly to mitigate the effects of possible outliers in
the data, we prefer to use the median values of specific ab-
sorption at T1 and T2 rather than mean values. Moreover,
given the skewed nature of the distributions of specific ab-
sorption shown in Fig. 9, we believe the median is a more
useful indicator of the behavior than the mean.
Histograms of specific absorption at the two sites are
shown in Fig. 9 and indicate that the distributions were
skewed toward higher values at T2 but not at T1. This is
what would be expected if the differences between the sites
are attributable to the greater aging of the EC sampled at the
site more distant from sources in Mexico City.
Fig. 8. Scatter plots of absorption as a function of EC concentration
at T1 (top) and T2 (bottom). The straight lines are least squares fits
to the data that are constrained to pass through the origin.
It is interesting to compare the values of specific absorp-
tion at the two sites during times when the Mexico City
plume was traveling from the city over T1 and then T2 with
the values found when such transport was not occurring. Us-
ing the T1 RWP data, we identified “transport” periods when
air parcels passing through T1 were likely to have originated
over the city and subsequently passed within 5km of T2 ap-
proximately 4h or less after exiting the city. In some cases
near-surface trajectories did not have favorable directions but
trajectories further aloft (e.g., 1 km) did. If these cases oc-
curred when a convective boundary layer was present that
could mix air parcels throughout the boundary layer to the
surface, then those periods were also included in the trans-
port category. All other periods were lumped into the “non-
transport” category. Without using a detailed mesoscale
model constrained by data assimilation of RWP data there is
some uncertainty and subjectivity in the choice of the partic-
ular time periods to include in each category, but our results
are not especially sensitive to the details of the selection cri-
teria. Table 1 gives median and 10th and 90th percentile val-
ues of the specific absorption at T1 and T2 for the transport
and non-transport periods. We have also included values of
www.atmos-chem-phys.net/7/1585/2007/ Atmos. Chem. Phys., 7, 1585–1598, 2007
1594 J. C. Doran et al.: The T1-T2 study
Fig. 9. Histograms of specific absorption of EC at 870 nm at T1
(top) and T2 (bottom). The histogram for T1 represents 264h of
data while that for T2 represents 236 h.
specific absorption extrapolated to a wavelength of 550nm,
assuming a λ1dependence for the absorption between 870
and 550nm as suggested by Bergstrom et al. (2002).
Two features stand out in the table. The first is that larger
values of specific absorption are found at both T1 and T2
during non-transport times than during transport times. This
is consistent with the expectation that longer-aged EC with
greater coating is more likely to be found at both sites if the
wind is not blowing directly from the city. The second fea-
ture is that the specific absorption at T2 is larger than at T1
for both transport and non-transport periods. This suggests
that the specific absorption of EC can be modified after only
a few hours of aging, the time required for transport from T1
to T2. On non-transport days the results indicate that fresher
emissions from Mexico City or nearby local sources are still
more likely to be found at T1 than at T2. Although the dif-
ferences in specific absorption noted here are not large, a
Mann-Whitney test indicates that the differences are statis-
tically significant at the 1% level.
The median values extrapolated to 550nm displayed
in the table are considerably larger than the value of
7.5±1.2m2g1reported by Bond and Bergstrom (2006)
for uncoated soot particles. However, Bond et al. (2006)
suggest that absorption by aged aerosol is about 1.5 times
greater than that of fresh aerosol at 550nm. Johnson et
al. (2005) note that while in Mexico City “fresh particulate
emissions from mixed-traffic are almost entirely carbona-
ceous...ambient soot particles which have been processed
for less than a few hours are heavily internally mixed...
Our larger values of specific absorption are thus consistent
with this characterization of the emissions.
It is interesting that the 10th percentile value of spe-
cific absorption at T1 during transport conditions is
8.7m2g1(extrapolated to 550nm), which is near the upper
end of the range of values suggested by Bond and Bergstrom
for freshly emitted soot. In contrast, the 10th percentile value
of specific absorption at T2 is 10.8, well beyond the sug-
gested range. For non-transport periods, the 10th percentile
value at T1 is still 8.7, which is consistent with our expec-
tation that even during non-transport periods T1 would often
be influenced by nearby sources. This is less likely to be
the case at T2, and the 10th percentile value there is indeed
substantially larger than for transport conditions.
The extrapolations of the specific absorption values from
870nm to 550nm shown in the table may be problematic.
Bergstrom et al. (2002) found a σAλ1dependence, where
σAis the cross section for aerosol scattering, but Moosm¨
uller
et al. (1998) derived a λ2.7relationship from their analysis.
Using PSAP data at three wavelengths we have examined
the wavelength dependence of absorption between 470 and
660nm. We found the median exponent in the expression
σAλbis close to 1, but there is considerable scatter about
this value. Moreover, an extrapolation derived from 660 and
530nm PSAP data cannot be applied with much confidence
to the much larger extrapolation of the 870nm photoacoustic
results to 550nm. Thus we do not believe that a point-by-
point comparison of 530nm PSAP and 870nm PAS absorp-
tion values is warranted. The extrapolation of the summary
statistics shown in the table may still be useful, but we sug-
gest that the 550 nm values be treated with considerable cau-
tion.
The G-1 aircraft did not measure EC so it is not possible to
use G-1 data to directly extract values of specific absorption
as it passed through the Mexico City urban plumes.
5.3 Single scattering albedo
We have made an initial examination of the single scatter-
ing albedo determined from the airborne nephelometer and
PSAP data to see if significant changes could be discerned
between the T1 and T2 locations on days when the urban
plume was carried over or near both sites. The single scat-
tering albedo is defined by $0=6S
6S+6A, where 6Sand 6A
are the extinction coefficients for aerosol scattering and ab-
sorption, respectively. Values of $0were calculated when
the G-1 was sampling within 5km of T1 or T2 on DOY 77,
78, and 79 (18, 19, and 20 March). From the PSAP and
Atmos. Chem. Phys., 7, 1585–1598, 2007 www.atmos-chem-phys.net/7/1585/2007/
J. C. Doran et al.: The T1-T2 study 1595
Table 1. Median, 10th and 90th percentile values of specific absorption of EC in m2g1for transport and non-transport periods.
T1 T2
10th percentile median 90th percentile 10th percentile median 90th percentile
transport peri-
ods 870 nm 5.5 7.7 9.2 6.8 8.3 11.9
non-transport
periods
870nm
5.5 8.1 9.7 7.5 9.0 12.2
transport peri-
ods 550 nm 8.7 12.2 14.5 10.8 13.1 18.8
non-transport
periods
550nm
8.7 12.8 15.4 11.8 14.3 19.3
Table 2. Mean values of $0derived from the T1 and T2 MFRSRs (at 500nm) and from the T2 surface aerosol system (nephelometer and
PSAP) data (at 530 nm) for DOY 71, 78, and 86 (12, 19, and 27 March). Note that surface values were available only at T2.
DOY Time (LST) $0from T1
MFRSR $0from T2 sur-
face system Time (LST) $0from
T2 MFRSR $0from T2 sur-
face system
71 08:00–09:36 0.84 0.88 08:00–09:36 0.91 0.88
78 07:30–09:30 0.85 0.92 07:30–11:28 0.83 0.91
86 08:00–10:00 0.89 0.92 07:30–08:00 0.90 0.90
nephelometer data, $0was on the order of 0.9 for most of
the flights during these transport periods, so that σAσS/9.
From Table 1, the median specific absorption at the surface
at T2 was 10% larger than that at T1. We can assume that
the variations in the surface absorption values were approxi-
mately the same as those at the G-1 sampling elevations be-
cause the boundary layer was generally well developed at the
times of the G-1 overflights. Thus, reductions in $0between
T1 and T2 arising from changes in σA(and assuming σSwas
unchanged) would only be on the order of 0.01. Given the
observed fluctuations in the values of $0, consistent differ-
ences between T1 and T2 were expected to be difficult to
identify in the G-1’s data stream (although such accuracy has
been suggested as needed for climate studies (Heintzenberg
et al., 1997)). This expectation was confirmed in our initial
examination of the G-1 data time series. Further analyses are
planned using CO measurements to help identify the times
and locations when the G-1 was sampling within the Mexico
City urban plume and to compare in-plume and out-of-plume
values of $0, but at this point we do not anticipate major
changes in $0will be found between T1 and T2.
We have evaluated single scattering albedos at the surface
at T2 for the transport and non-transport periods described
above as well. The distributions of $0for these two peri-
ods are similar, with mean values differing by approximately
0.01 or less. This implies that the direct radiative forcing
at T2 should not be significantly affected by the differences
in the specific absorption of EC between transport and non-
transport periods, although it still could be affected by other
factors such as variations in the total optical depth between
the two sites. It would have been desirable to compare sur-
face results from the T1 and T2 sites to see if those dis-
tributions of $0were similar, but the data recovery for the
nephleometer and PSAP at T1 was poor and we are unable
to provide such a comparison at this time.
We have also used the MFRSR data at T1 and T2 to de-
rive values for $0at 500nm using the approach described by
Kassianov et al. (2005). In contrast to the aerosol quantities
measured by our other surface instruments, $0derived from
the MFRSR data can be considered as an average over the
atmospheric column. The approach is only applicable when
the sky is essentially cloud-free. In practice this often limits
our analysis periods to a few hours in the morning because
cumulus clouds frequently developed in the late morning or
afternoon. Initial results are similar to those obtained from
the G-1 measurements described above, i.e., differences in
$0between T1 and T2 are small (0.01–0.02) without an
obvious tendency for values at one or the other site to be
larger.
Table 2 shows some values of $0for a wavelength of
500nm obtained from the MFRSRs and from the T2 sur-
face aerosol system at 530nm on three days with cloud-free
www.atmos-chem-phys.net/7/1585/2007/ Atmos. Chem. Phys., 7, 1585–1598, 2007
1596 J. C. Doran et al.: The T1-T2 study
periods. The times of the MFRSR measurements are given in
the table. The agreement between the column averaged val-
ues of $0and the T2 surface value is good for DOY 86 (27
March), fair for DOY 71, and poorer on DOY 78 (19 March).
The discrepancy between MFRSR and surface values of
$0on DOY 78 may be due to boundary-layer structure and
growth. The G-1 made measurements on that day between
10:50 and 12:00 LST and found values ranging from 0.87
to 0.90 over T1 and around 0.87 over T2. At T2 this re-
sult is an intermediate one between the surface value and
the column-averaged value. The sampling height of the air-
craft was 3100m m.s.l., or approximately 830 m a.g.l. at
T1 and 560m a.g.l. at T2. This would have placed the G-1
within the growing boundary layer at heights of 0.5–0.7 of
the mixed layer depths estimated from RWP signal-to-noise
values. It is thus reasonable to assume that the G-1 was sam-
pling in a region where the aerosol mix was influenced by
both near-surface aerosols, with higher values of $0as in-
dicated by the nephelometer and PSAP data, and aerosols
entrained from aloft, with lower values of $0as suggested
by the MFRSR results. Later in the afternoon (17:00 LST)
when the boundary layer was deeper, the G-1 was again over
the T2 site and measured a value of 0.90 for $0, which agrees
well with the corresponding surface value at that time of
0.92.
6 Summary
We have carried out an experiment to investigate the evolu-
tion of aerosols and their optical properties downwind from
Mexico City. Our focus has been on the specific absorption
of black carbon and how that is affected as the aerosols age
and become coated, changing from an externally to an inter-
nally mixed state. To accomplish this we deployed instru-
ments at two sites, T1 and T2, to characterize the meteorol-
ogy, solar radiation, aerosol light scattering and absorption,
and OC and EC concentrations. An instrumented aircraft
flew a number of missions over the two sites as well.
The purpose of this paper has been to provide an overview
of the T1-T2 experiment and to present some preliminary re-
sults and highlights from that campaign. OC and EC values
at T1 showed diurnal variations suggesting a buildup of more
polluted air from nearby urban sources during the night and a
subsequent dilution as the boundary layer grew the following
morning. This behavior was not found at the more isolated
T2 site but transport from Mexico City appeared to be a ma-
jor factor in determining OC and EC concentrations. Using
data from a radar wind profiler, we have divided the mea-
surement periods into transport and non-transport periods,
corresponding to conditions when air masses were or were
not expected to travel from Mexico City and then over T1
and T2. The specific absorption of EC during transport peri-
ods was lower than that found during non-transport periods,
which is consistent with the expectation that fresher emis-
sions are more likely to be found at T1 and T2 during trans-
port episodes than at other times. The specific absorption at
T2, the site more distant from Mexico City, was also found
to be larger than at T1, in keeping with the greater age of the
aerosols expected at T2. Initial analyses of the variation of
$0at T2 for transport and non-transport periods reveal only
small differences, although additional analyses are planned.
Comparisons of values from G-1 aircraft data, surface opti-
cal measurements, and column-averaged values derived from
MFRSR data appear broadly consistent, although boundary-
layer structure and growth appear to contribute to some dif-
ferences between the in situ and the column-averaged values.
A number of additional topics are anticipated for future
studies, including detailed examination of the in-plume and
out-of-plume aerosol characteristics derived from the G-1
measurements, more detailed analyses of plume trajectories
and their relation to the specific absorption measured at T1
and T2, assessment of the aerosol optical characteristics at
T0, and further comparisons of in situ and column-averaged
aerosol properties.
Acknowledgements. We thank J. Ogren, B. Andrews, and P. Sheri-
dan of NOAA’s CMDL for their assistance with the construction
of the aerosol optical measurement systems and the subsequent
data processing. We appreciate the assistance of P. Holowecky
and J. Satola of Battelle Columbus for their contributions to the
field measurements. This research was supported by the Office of
Science (BER), U.S. Department of Energy, under the auspices
of the Atmospheric Science Program, under Contract DE-AC05-
76RL01830 at the Pacific Northwest National Laboratory. Pacific
Northwest National Laboratory is operated for the U.S. DOE by
Battelle Memorial Institute.
Edited by: S. Madronich
References
Anderson, T. L. and Ogren, J. A.: Determining aerosol radiative
properties using the TSI 3563 Integrating Nephelometer, Aerosol
Sci. Technol., 29, 57–69, 1998.
Andreae, M. O. and Gelencs´
er, A.: Black carbon or brown carbon?
The nature of light-absorbing aerosols, Atmos. Chem. Phys., 6,
3131–3148, 2006,
http://www.atmos-chem-phys.net/6/3131/2006/.
Arnott, W. P., Moosm¨
uller, H., Sheridan, P. J., Ogren, J. A.,
Raspert, R., Slaton, W. V., Hand, J. L., Kreidenweis, S. M.,
and Collett Jr., J. L.: Photoacoustic and filter-based ambient
aerosol light absorption measurements: instrument comparisons
and the role of relative humidity, J. Geophys. Res., 108, 4034,
doi:10.1029/2002JD002165, 2003.
Baumgardner, D., Raga, G. B., Kok, G., Ogren, J., Rosas, I., B´
aez,
A., and Novakov, T.: On the evolution of aerosol properties at a
mountain site above Mexico City, J. Geophys. Res., 105, 22 243–
22 253, 2000.
Baumgardner, D. and Clarke, A.: Changes in aerosol properties
with relative humidity in the remote southern hemisphere ma-
Atmos. Chem. Phys., 7, 1585–1598, 2007 www.atmos-chem-phys.net/7/1585/2007/
J. C. Doran et al.: The T1-T2 study 1597
rine boundary layer, J. Geophys. Res., 103(D13), 16525–16 534,
doi:10.1029/98JD00688, 1998.
Baumgardner, D., Raga, G., Peralta, O., Rosas, I., Castro,
T., Kuhlbusch, T., John, A., and Petzold, A.: Diagnos-
ing black carbon trends in large urban areas using carbon
monoxide measurements, J. Geophys. Res., 107(D21), 8342,
doi:10.1029/2001JD000626, 2002.
Birch, M. E. and Cary, R. A.: Elemental carbon-based method for
monitoring occupational exposures to particulate diesel exhaust,
Aerosol Sci. Technol., 25, 221–241, 1996.
Bond, T. C., Anderson, T. L., and Campbell, D.: Calibration and
intercomparison of filter-based measurements of visible light ab-
sorption by aerosols, Aerosol Sci. Technol., 30, 582–600, 1999.
Bond, T. C. and Bergstrom, R. W.: Light absorption by carbona-
ceous particles: an investigative review. Aerosol Sci. Technol.,
40, 27–67, doi:10.1080/02786820500421521, 2006.
Bond., T. C., Habib, G., and Bergstrom, R. W.: Limitations in the
enhancement of visible light absorption due to mixing state, J.
Geophys. Res., 111, D20211, doi:10.1029/2006JD007315, 2006.
Carrico, C. M., Kus, P., Rood, M. J., Quinn, P. K., and Bates,
T. S.: Mixtures of pollution, dust, sea salt, and volcanic
aerosol during ACE-Asia: Radiative properties as a func-
tion of relative humidity, J. Geophys. Res., 108(D23), 8650,
doi:10.1029/2003JD003405, 2003.
Chung, S. H. and Seinfeld, J. H.: Global distribution and climate
forcing of carbonaceous aerosols, J. Geophys. Res., 107(D19),
4407, doi:10.1029/2001JD001397, 2002.
Doran, J. C., Abbott, J. L., Archuleta, J., Bian, X., Chow, J. C.,
Coulter, R. L., de Wekker, S. F. J., Edgerton, S. A., Fernandez,
A., Fast, J. D., Hubbe, J. M., King, C. W., Langley, D., Leach, J.
M., Lee, J. T., Martin, T. J., Martinez, D., Martinez, J. L., Mer-
cado, G., Mora, V., Mulhearn, M., Pena, J. L., Petty, R., Porch,
W. M., Russell, C., Salas, R., Shannon, J. D., Shaw, W. J., Sosa,
G., Watson, J. G., Templeman, B., White, R., Whiteman, C. D.,
and Wolfe, D.: The IMADA-AVER boundary layer experiment
in the Mexico City area, Bull. Am. Meteorol. Soc., 79, 2497–
2508, 1998.
Fuller, K. A., Malm, W. C., and Kreidenweis, S. M.: Effects of mix-
ing on extinction by carbonaceous particles, J. Geophys., Res.,
104, 15 941–15954, 1999.
Harrison, L., Michalsky, J., and Berndt, J.: Automated multifil-
ter rotating shadow-band radiometer: an instrument for optical
depth and radiation measurements, Appl. Opt., 33, 5118–5125,
1994.
Heintzenberg, J., Charlson, R. J., Clarke, A. D., Liousse, C.,
Ramaswamy, V., Shine, K. P., Wendisch, M., and Helas, G.:
Measurements and modeling of aerosol single-scattering albedo:
Progress, problems, and prospects, Beitr. Phys. Atmosph., 70,
249–263, 1997.
Horvath, H.: Atmospheric light absorption – A review, Atmos. En-
viron., 27A(3), 293–317, 1993.
Im, J.-S., Saxena, V. K., and Wenny, B. N.: An assessment of hy-
groscopic growth factors for aerosols in the surface boundary
layer for computing direct radiative forcing, J. Geophys. Res.,
106(D17), 20 213–20224, doi:10.1029/2000JD000152, 2001.
Jacobson, M. Z.: Isolating nitrated and aromatic aerosols and ni-
trated aromatic gases as sources of ultraviolet light absorption, J.
Geophys. Res.-A, 104(D3), 3527–3542, 1999.
Jacobson, M. Z.:. Strong radiative heating due to the mixing state
of black carbon in atmospheric aerosols, Nature, 409, 695–696,
2001.
Jacobson, M. Z. and Seinfeld, J. H.: Evolution of nanoparticle size
and mixing state near the point of emission, Atmos. Environ., 38,
1839–1850, 2004.
Johnson, K. S., Zuberi, B., Molina, L. T., Molina, M. J., Iedema,
M. J., Cowin, J. P., Gaspar, D. J., Wang, C., and Laskin, A.:
Processing of soot in an urban environment: case study from the
Mexico City Metropolitan Area, Atmos. Chem. Phys., 5, 3033–
3043, 2005,
http://www.atmos-chem-phys.net/5/3033/2005/.
Kassianov, E., Barnard, J., and Ackerman, T. P.: Retrieval of aerosol
microphysical properties using surface multifilter rotating shad-
owband radiometer (MFRSR) data: modeling and observations,
J. Geophys. Res., 110, D09201, doi:10.1029/2004JD005337,
2005.
Kirchstetter, T. W., Novakov, T., and Hobbs, P. V.: Evidence that
the spectral dependence of light absorption by aerosols is af-
fected by organic carbon, J. Geophys. Res.-A, 109, D21208,
doi:10.1029/2004JD004999, 2004.
Liousse, C., Cachier, H., and Jennings, S. G.: Optical and thermal
measurements of black carbon aerosol content in different en-
vironments: variation of the specific attenuation cross-section,
sigma (σ ), Atmos. Environ., 27A, 1203–1211, 1993.
Mallet, M., Roger, J. C., Despiau, S., Dubovik, O., and Putaud, J.
P.: Microphysical and optical properties of aerosol particles in
urban zone during ESCOMPTE, Atmos. Res., 69, 73–97, 2003.
Markowicz, K. M., Flatau, P. J., Quinn, P. K., Carrico, C.M., Flatau,
M. K., Vogelmann, A. M., Bates, D., Liu, M., and Rood, M. J.:
Influence of relative humidity on aerosol radiative forcing: An
ACE-Asia experiment perspective, J. Geophys. Res., 108(D23),
8662, doi:10.1029/2002JD003066, 2003.
Marley, N.A., Gaffney, J. S., Baird, C., Blazer, C. A., Drayton, P. J.,
and Frederick, J. E.: An empirical method for the determination
of the complex refractive index of size-fractionated atmospheric
aerosols for radiative transfer calculations, Aerosol Sci. Technol.,
34(6), 535–549, 2001.
Mikhailov, E. F., Vlasenko, S. S., Podgorny, I. A., Ramanathan,
V., and Corrigan, C. E.: Optical properties of soot-water drop
agglomerates: an experimental study, J. Geophys. Res., 111,
D07209, doi:10.1029/2005JD006389, 2006.
Moosm¨
uller, H., Arnott, W. P., Rogers, C. F., Chow, J. C., Frazier,
C. A., Sherman, L. E., and Dietrich, D. L.: Photoacoustic and
filter measurements related to aerosol light absorption during the
Northern Front Range Air Quality Study (Colorado 1996/1997),
J. Geophys. Res., 103(D21), 28149–28 157, 1998.
Morse, C. S., Goodrich, R. K., and Cornman, L. B.: The NIMA
method for improved moment estimation from Doppler spectra,
J. Atmos. Oceanic Technol., 19, 274–295, 2002.
Nemesure, S., Wagener, R., and Schwartz, S. E.: Direct shortwave
forcing of climate by the anthropogenic sulfate aerosol: sensitiv-
ity to particle size, composition, and relative humidity, J. Geo-
phys. Res., 100(D12), 26105–26 116, doi:10.1029/95JD02897,
1995.
Penner, J. E, Chang, C. C., and Grant, K.: Climate forcing by car-
bonaceous and sulfate aerosols, Clim. Dyn., 14, 839–851, 1998.
Petzold, A., Kopp, C., and Niessner, R.: The dependence of the
specific attenuation cross-section on black carbon mass fraction
and particle size, Atmos. Environ., 35(5), 661–672, 1997.
www.atmos-chem-phys.net/7/1585/2007/ Atmos. Chem. Phys., 7, 1585–1598, 2007
1598 J. C. Doran et al.: The T1-T2 study
Redemann, J., Russell, P. B., and Hamill, P.: Dependence
of aerosol light absorption and single-scattering albedo on
ambient relative humidity for sulfate aerosols with black
carbon cores, J. Geophys. Res., 106(D21), 27 487–27 496,
doi:10.1029/2001JD900231, 2001.
Sato, M., Hansen, J., Koch, D., Lacis, A., Ruedy, R., Dubovik,
O., Holben, B., Chin, M., and Novakov, T.: Global atmospheric
black carbon inferred from AERONET, Proc. Natl. Acad. Sci.,
100, 6319–6324, 2003.
Schnaiter, M., Horvath, H., M¨
ohler, O., Naumann, K.-H., Saathoff,
H., and Sch¨
ock, O. W.: UV-VIS-NIR spectral optical properties
of soot and soot-containing aerosols, J. Aerosol Sci., 34, 1421–
1444, 2003.
Schuster, G. L., Dubovik, O., Holben, B. N., and Clothiaux,
E. C.: Inferring black carbon content and specific absorption
from Aerosol Robotic Network (AERONET) aerosol retrievals,
J. Geophys. Res., 110, D10S17, doi:10.1029/2004JD004548,
2005.
Sokolik, I. N. and Toon, O. B.: Incorporation of mineralogical com-
position of aerosols into models of radiative properties of min-
eral aerosol from the UV to IR wavelengths, J. Geophys. Res.-A,
104(D8), 9423–9444, 1999.
Volkamer, R., Jimenez, J. L., San Martini, F., Dzepina, K., Zhang,
Q., Salcedo, D., Molina, L. T., Worsnop, D. R., and Molina, M.
J.: Secondary organic aerosol Formation from anthropogenic air
pollution: rapid and higher than expected, Geophys. Res. Lett.,
33, L17811, doi:10.1029/2006GL026899, 2006.
Waggoner, A. P., Weiss, R. E., Ahlquist, N. C., Covert, D. S., Will,
S., and Charlson, R. J.: Optical characteristics of atmospheric
aerosols, Atmos. Environ., 15, 1891–1909, 1981.
Atmos. Chem. Phys., 7, 1585–1598, 2007 www.atmos-chem-phys.net/7/1585/2007/
... Mexico City, one of the world's most populated urban areas, experiences a heavy daily load of automobile and industrial emissions, which strongly degrade air quality (Molina et al. 2007;Chavez-Baeza and Sheinbaum-Pardo 2014;Peralta et al. 2019). Moreover, the topographically complex terrain, in which this city is located, results in an complex interaction between synoptic and local circulations forced by the mountainous terrain (Doran et al. 1998;Jáuregui 1988;Whiteman et al. 2000;Jazcilevich et al. 2003;Doran 2007;Foy et al. 2005Foy et al. , 2008Molina et al. 2007;Pozo et al. 2019;Díaz-Esteban et al. 2022). This complex circulation can stifle ventilation and the dispersion of pollutants (Jazcilevich et al. 2003;Doran 2007;Foy et al. 2006;Díaz-Esteban et al. 2022;García-Franco 2020;Burgos-Cuevas et al. 2021). ...
... Moreover, the topographically complex terrain, in which this city is located, results in an complex interaction between synoptic and local circulations forced by the mountainous terrain (Doran et al. 1998;Jáuregui 1988;Whiteman et al. 2000;Jazcilevich et al. 2003;Doran 2007;Foy et al. 2005Foy et al. , 2008Molina et al. 2007;Pozo et al. 2019;Díaz-Esteban et al. 2022). This complex circulation can stifle ventilation and the dispersion of pollutants (Jazcilevich et al. 2003;Doran 2007;Foy et al. 2006;Díaz-Esteban et al. 2022;García-Franco 2020;Burgos-Cuevas et al. 2021). The characterization of the ABL and its diurnal evolution in this urbanized zone are crucial for assessing the vertical dispersion of pollutants. ...
Article
Full-text available
The Atmospheric Boundary Layer (ABL) height is a key parameter in air quality research as well as for numerical simulations and forecasts. The identification of thermally stable layers, often with radiosondes, has been a common approach for estimating ABL height, though with limited temporal coverage. Remote sensing techniques offer essentially continuous measurements. Nevertheless, ABL height retrievals from different methods can vary greatly when compared, which is particularly notable for topographically complex terrains, such as that surrounding Mexico City. This study, employing one year of data in Mexico City, reveals that the daytime convective boundary layer height (retrieved from Doppler lidar data) is typically lower than the aerosol layer height (retrieved from ceilometer data). Although both estimated heights evolved diurnally, the more elevated aerosol layer decays more slowly, suggesting that the mechanisms that elevate aerosols are not limited to convective motions. Additionally, both diurnal and seasonal variability are investigated, comparing remotely sensed-retrieved heights with thermally stable layers estimated from radiosonde data. Multiple stable layers often develop, those at higher levels have similar values to the ceilometer-retrieved heights, while stable layers at lower heights are similar to Doppler lidar height retrievals. The present research constitutes the first detailed analysis of ceilometer backscatter and Doppler lidar thresholding methods for estimating ABL height over Mexico City, and our results illustrate the complexity of mixing mechanisms on the ABL in this region of complex orography.
... The atmosphere of MC has been widely studied and, like many urban atmospheres, contains high concentrations of PM with different mixtures of organic and inorganic compounds. The rate of SOA formation is fast, so internal mixing occurs within a few hours after emission [27,[29][30][31][32]. Many studies on the optical properties of aerosols in MC focus on light absorption by BC or BrC. ...
Article
Full-text available
From January to March 2015, an atmospheric aerosol measurement campaign, “Aerosoles en Ciudad Universitaria 2015” (ACU15), was carried out in Mexico City to determine the particles’ optical properties and chemical composition. Two photoacoustic spectrometers measured the scattering and absorption coefficient at two different wavelengths. The average absorption coefficient at 532 nm was 12.71 ± 9.48 Mm−1 and at 870 nm was 10.35 ± 7.36 Mm−1. The average scattering coefficient was 65.63 ± 47.12 Mm−1 (532 nm) and 21.12 ± 14.24 Mm−1 (870 nm). The chemical composition was determined via an aerosol chemical speciation monitor. The organic aerosol fraction represented 53% of the total PM1 and was made up of 63% low volatile (4.64 µg m−3), 22% hydrogenated (1.90 µg m−3), and 15% semi-volatile organics (1.54 µg m−3). The correlation coefficient of chemical species (NO3−, NH4+, SO42−, low-volatile, and semi-volatile organics) and optical properties was 0.92. The multilinear regression showed a good agreement among chemical species and optical properties (r > 0.7). The mass absorption coefficient calculated for the measuring site at 870 nm was MAE870 = 5.8 m2 g−1, instead of the default 4.74 m2 g−1. Furthermore, based on the median AAE, the 532 nm MAE532 resulting from the multiple linear regression (MLR) showed the following coefficients: 7.70 m2 g−1 (eBC), 0.22 m2 g−1 (HOA), and 0.16 m2 g−1 (LV–OOA). The coefficients of MLR were: 7.08 m2 g−1 (eBC), 5.83 m2 g−1 (NO3−), 5.69 m2 g−1 (low volatile organic aerosol), 2.78 m2 g−1 (SO42−), 2.40 m2 g−1 (hydrocarbon-like organic aerosol), and 1.04 m2 g−1 (semi volatile organic aerosol).
... The NEXAFS spectrum for type 2 is quantitatively similar to those reported for organic particles from anthropogenic emissions in urban areas of of Mexico City (Moffet et al., 2010b) and Central California . As the aerosol plume is transported away from the source of emission, organic mass increases while the fraction of C=C decreases (Doran et al., 2007;Kleinman et al., 2008;Moffet et al., 370 2010b). As a result, organic functional groups build up with particle age such as carboxylic acids, carbonyl, alcohol, and other carbon-oxygen functional groups. ...
Preprint
Full-text available
Long-range transport of continental emission has far reaching influence over remote regions resulting in substantial change in the size, morphology, and composition of the local aerosol population and cloud condensation nuclei (CCN) budget. Here, we investigate the physiochemical properties of atmospheric particles collected onboard a research aircraft flown over the Azores during the winter 2018 Aerosol and Cloud Experiment in the Eastern North Atlantic (ACE-ENA) campaign. Particles were collected within the marine boundary layer (MBL) and free troposphere (FT), after long-range atmospheric transport episodes facilitated by dry intrusion (DI) events. Chemical and physical properties of individual particles were investigated using complementary capabilities of computer-controlled scanning electron microscopy and X-ray spectro-microscopy to probe particle external and internal mixing state characteristics in the context of real-time measurements of aerosol size distribution, cloud condensation nuclei (CCN) concentration, and back trajectory calculations. While carbonaceous particles were found to be the dominant particle-type in the region, changes in the percent contribution of organics across the particle population (i.e., external mixing) shifted from 68 % to 43 % in the MBL and from 92 % to 46 % in FT samples during DI events. This change in carbonaceous contribution is counterbalanced by the increase of inorganics from 32 % to 57 % in the MBL and 8 % to 55 % in FT. The quantification of organic volume fraction (OVF) of individual particles derived from X-ray spectro-microscopy, which relates to the multi-component internal composition of individual particles, showed a factor of 2.06 ± 0.16 and 1.11 ± 0.04 increase in the MBL and FT, respectively, among DI samples. We show that supplying particle OVF into the κ-Köhler equation can be used as a good approximation of field measured in-situ CCN concentrations. We also report changes in the κ values between κMBL, non-DI = 0.48 to κMBL, DI = 0.41 and κFT, non-DI = 0.36 to κFT, DI = 0.33, which is consistent with enhancements in OVF followed by the DI episodes. Our observations suggest that entrainment of particles from long-range continental sources alters the mixing state population and CCN properties of aerosol in the region. The work presented here provides field observation data that can inform atmospheric models that simulate sources and particle composition in the Eastern North Atlantic.
... As one of the world's most populous cities, Mexico City experiences a heavy daily load of emissions from more than four million vehicles, as well as from industry, significantly affecting air quality (Molina et al. 2007;Chavez-Baeza and Sheinbaum-Pardo 2014;Peralta et al. 2019). Contamination transport resulting from local circulations in Mexico City and associated meteorological conditions has been investigated (Doran et al. 1998;Jáuregui 1988;Whiteman et al. 2000;Doran 2007;Molina et al. 2007). Results from these campaign studies have indicated that a close relationship exists between high-pollution episodes and thermal inversions in Mexico City. ...
Article
Full-text available
Lower tropospheric thermal structure greatly affects atmospheric boundary-layer (ABL) stability and mixing processes with the free troposphere. In particular, in polluted urban zones, ABL stratification becomes a key variable in air quality research. This study focuses on generating a climatology (1990-2017) of the seasonal variability of ABL thermal structure in Mexico City by way of radiosonde analysis. Thermal inversion intensity and frequency are shown to be greater during winter and spring, a behaviour which coincides with greater pollutant concentrations. Higher concentrations are found during the dry season (November to May) than during the rainy months. In addition, significantly higher than normal surface pollutant concentrations are found on days with simple thermal inversion layers as well as during multilayer inversion days. Furthermore, stable layers, determined by potential temperature , are found throughout the year but more frequently during winter, whereas stable layers based on the virtual potential temperature prevail all year. In regions of complex terrain, such multiple stable layers have also been identified by previous authors. Additionally, the most unstable surface layers (in which the bulk Richardson number (Ri B) is small) develop during the rainy season, whereas during winter there are more levels in the vertical column with higher Ri B values. Although the Mexico City ABL and pollution episodes have been widely studied, this represents the first long-term investigation to consider the thermal stability of the ABL. Therefore, the present study provides a baseline for further research employing different observational techniques and high-resolution numerical models.
Article
Full-text available
Brown carbon (BrC) is a component of particulate matter which has significant impacts on climate forcing and air quality. To elucidate the current state and sources of BrC in Mexico City, the largest megacity in North America, the aerosols' ability to scatter and absorb light was evaluated over two years in concert with detailed chemical speciation of their components during a period of six months. These measurements made it possible to evaluate the seasonal and diurnal variations of the chemical composition, optical properties, and origin of BrC. It was found that 67% of the light extinction at 870 nm is due to scattering. Organic aerosols dominate the submicron mass loading (62%), as well as the light scattering (>50%). Nitrate and sulfate compounds are also important contributors to light scattering. Among the organic fraction, fresh particles strongly associated with traffic emissions dominate the light absorption at ultraviolet (UV) wavelengths on days not affected by biomass burning plumes. Regional wildfires are ubiquitous in the central region of Mexico during the dry-warm season (March–May) and can drastically increase the light scattering and absorption attributed to the organic fraction. During wildfire episodes, the organic fraction can contribute up to 80% and 50% to light scattering and absorption, respectively. Aged organic aerosols have a negligible contribution to light absorption at UV wavelengths, but secondary organic aerosols of recent formation contribute on average 24% in days not affected by wildfire plumes. Brown carbon and black carbon (BC) contributed 22% and 78% to the total light absorption in México City, respectively. Brown carbon increases on average 28% the light absorption over that attributed to BC. This increase can be up to 32% during the dry-warm season. In summary, vehicular traffic is the main contributor to BrC light absorption on a daily basis, while biomass burning becomes the major contributor during wildfire episodes.
Article
Full-text available
Long-range transport of continental emissions has a far-reaching influence over remote regions, resulting in substantial change in the size, morphology, and composition of the local aerosol population and cloud condensation nuclei (CCN) budget. Here, we investigate the physicochemical properties of atmospheric particles collected on board a research aircraft flown over the Azores during the winter 2018 Aerosol and Cloud Experiment in the Eastern North Atlantic (ACE-ENA) campaign. Particles were collected within the marine boundary layer (MBL) and free troposphere (FT) after long-range atmospheric transport episodes facilitated by dry intrusion (DI) events. Chemical and physical properties of individual particles were investigated using complementary capabilities of computer-controlled scanning electron microscopy and X-ray spectromicroscopy to probe particle external and internal mixing state characteristics. Furthermore, real-time measurements of aerosol size distribution, cloud condensation nuclei (CCN) concentration, and back-trajectory calculations were utilized to help bring into context the findings from offline spectromicroscopy analysis. While carbonaceous particles were found to be the dominant particle type in the region, changes in the percent contribution of organics across the particle population (i.e., external mixing) shifted from 68 % to 43 % in the MBL and from 92 % to 46 % in FT samples during DI events. This change in carbonaceous contribution is counterbalanced by the increase in inorganics from 32 % to 57 % in the MBL and 8 % to 55 % in FT. The quantification of the organic volume fraction (OVF) of individual particles derived from X-ray spectromicroscopy, which relates to the multi-component internal composition of individual particles, showed a factor of 2.06 ± 0.16 and 1.11 ± 0.04 increase in the MBL and FT, respectively, among DI samples. We show that supplying particle OVF into the κ-Köhler equation can be used as a good approximation of field-measured in situ CCN concentrations. We also report changes in the κ values in the MBL from κMBL, non-DI=0.48 to κMBL, DI=0.41, while changes in the FT result in κFT, non-DI=0.36 to κFT, DI=0.33, which is consistent with enhancements in OVF followed by the DI episodes. Our observations suggest that entrainment of particles from long-range continental sources alters the mixing state population and CCN properties of aerosol in the region. The work presented here provides field observation data that can inform atmospheric models that simulate sources and particle composition in the eastern North Atlantic.
Article
Full-text available
The urban climatology consists not only of the urban canopy temperature but also of wind regime and boundary layer evolution among other secondary variables. The energetic input and response of urbanized areas is rather different to rural or forest areas. In this paper, we outline the physical characteristics of the urban canopy that make its energy balance depart from that of vegetated areas and change local climatology. Among the several canopy characteristics, we focus on the aspect ratio h/d and its effects. The literature and methods of retrieving meteorological quantities in urban areas are reviewed and a number of physical analyzes from conceptual or numerical models are presented. In particular, the existence of a maximum value for the urban heat island intensity is discussed comprehensively. Changes in the local flow and boundary layer evolution due to urbanization are also discussed. The presence of vegetation and water bodies in urban areas are reviewed. The main conclusions are as follows: for increasing h/d, the urban heat island intensity is likely to attain a peak around h/d approximate 4 and decrease for h/d > 4; the temperature at the pedestrian level follows similar behavior; the urban boundary layer grows slowly, which in combination with low wind, can worsen pollution dispersion.
Article
Full-text available
Black carbon (BC), carbon monoxide (CO), and carbon dioxide (CO2) were measured in Mexico City (UNAM observatory) with a Photoacoustic Extinctiometer-PAX (BC) and a Cavity Ring-Down Spectroscopy analyzer-CRDS (CO and CO2), from November 2014 to July 2016. The objective of this study was to determine temporal variations of BC, CO, and CO2, their mutual correlations, and evaluation of the Mexico City emission inventory. The highest concentrations of pollutants were detected in cold dry season. The average concentrations of BC, CO, and CO2, for the entire period, were 2.95 μg m⁻³, 0.64 ppm, and 421.81 ppm, respectively. We calculated ΔBC/ΔCO, ΔBC/ΔCO2, and ΔCO/ΔCO2 using three methods to obtain a confidence interval for the emission ratios. BC, CO, CO2 concentrations, and the ΔBC/ΔCO ratio were maximum in the early morning, while the ΔBC/ΔCO2 and ΔCO/ΔCO2 peak was maximum in the afternoon. BC and CO have a weekday/weekend difference. The estimated slopes (ΔBC/ΔCO, ΔBC/ΔCO2, and ΔCO/ΔCO2) were compared with the emission ratios (BC/CO, BC/CO2, and CO/CO2) derived of the 2016 Mexico City emissions inventory. For mobile sources, the emission ratios of BC/CO and BC/CO2 were within the measurement range of ΔBC/ΔCO and ΔBC/ΔCO2, while CO/CO2 was underestimated in the emission inventory.
Article
The light absorption black carbon (BC) and brown carbon (BrC) are two important sources of uncertainties in radiative forcing estimate. Here we investigated the light absorption enhancement of (Eabs) of BC due to coated materials at an urban (Beijing) and a rural site (Gucheng) in North China Plain (NCP) in winter 2019 by using a photoacoustic extinctiometer coupled with a thermodenuder. Our results showed that the average (±1σ) Eabs was 1.32 (±0.15) at the rural site, which was slightly higher than that at the urban site (1.24±0.15). The dependence of Eabs on coating materials was found to be relatively limited at both sites, however, Eabs presented considerable increases as a function of relative humidity below 70%. Further analysis showed that Eabs during non-heating period in Beijing was mainly caused by secondary components, while it was dominantly contributed by enhanced primary emissions in heating season at both sites. In particular, aerosol particles mixed with coal combustion emissions had a large impact on Eabs (>1.40), while the fresh traffic emissions and freshly oxidized secondary OA (SOA) had limited Eabs (1.00~1.23). Although highly aged or aqueous-phase processed SOA coated on BC showed the largest Eabs, their contributions to the bulk absorption enhancement were generally small. We also quantified the absorption of BrC and source contributions. The results showed the BrC absorption at the rural site was nearly twice that of urban site, yet absorption Ångström exponents were similar. Multiple linear regression analysis highlighted the major sources of BrC being coal combustion emissions and photochemical SOA at both sites with additional biomass burning at the rural site. Overall, our results demonstrated the relatively limited winter light absorption enhancement of BC in different chemical environments in NCP, which needs be considered in regional climate models to improve BC radiative forcing estimates.
Article
Full-text available
Absorption by light-absorbing carbon (LAC) particles increases when the carbon is mixed with other material, and this change affects climate forcing. We investigate this increase theoretically over a realistic range of particle sizes. Perfect mixing at the molecular level often overestimates absorption. Assuming that LAC is coated by a concentric shell of weakly absorbing material, we calculate absorption by a range of realistic particle sizes and identify regimes in which absorption behaves similarly. We provide fits to amplification in five regions: (1) small cores and (2) intermediate cores, both with large shells; (3) small to intermediate cores with intermediate shells; (4) cores with growing shells; and (5) intermediate to large cores with large shells. Amplification in region 1 is highest but is physically implausible. Amplification in region 5 is constant at about 1.9 and represents an asymptote for particles with broad size distributions. Because absorption by aggregates is amplified by about 1.3 above spherical particles, and that factor is lost when particles are coated, we suggest that absorption by aged aerosol is about 1.5 times greater than that of fresh aerosol. The rate at which particles acquire sufficient coating to increase their original diameter by 60% is important in determining total absorption during their atmospheric lifetimes. Fitted amplification factors are not very sensitive to assumed refractive index of LAC and can be used even in simple models.
Article
Full-text available
Size distributions, scattering and absorption coefficients, and the bulk chemical composition of aerosols have been measured at a mountain site 400 m above the southwest sector of the Mexico City basin during a two-week period in November 1997. Variations in these properties are primarily related to local meteorology, i.e., wind direction and relative humidity; however, a link was found between carbon monoxide and ozone and the partitioning of aerosols between Aitken and accumulation mode sizes. Relative humidity was also found to affect this partitioning of aerosol size and volume. In addition, the fraction of sulfate in the aerosols was much higher on a high-humidity day than on a very low humidity day; however, the fraction of the mass contained in organic and elemental carbon was the same regardless of humidity levels. The daily variations of aerosol properties are associated with the arrival of new particles at the research site transported from the city basin and their subsequent mixture with aged aerosols that remain in the residual layer from the night before.
Article
Full-text available
Measurements in 1973 and 1987 showed that downward ultraviolet (UV) irradiances within the boundary layer in Los Angeles were up to 50% less than those above the boundary layer. Downward total solar irradiances were reduced by less than 14% in both studies. It is estimated that standard gas and particulate absorbers and scatterers accounted for only about 52–62% of the observed UV reductions at Claremont and Riverside. It is hypothesized that absorption by nitrated and aromatic aerosol components and nitrated aromatic gases caused at least 25–30% of the reductions (with aerosols accounting for about 4/5 of this percent). The remaining reductions are still unaccounted for. Absorbing aerosol components include nitrated aromatics, benzaldehydes, benzoic acids, aromatic polycarboxylic acids, phenols, polycyclic aromatic hydrocarbons, and nitrated inorganics. Many of these species have been observed to date in atmospheric particles, and absorption coefficient data indicate many are strong absorbers at long UV wavelengths. Since aerosols containing nitrated or aromatic aerosols have been observed widely in many areas aside from Los Angeles the finding may account for a portion of UV extinction in those regions as well. In Los Angeles, the finding may be important for predicting smog evolution, since UV reductions associated with high aerosol loadings were estimated to cause a 5–8% decrease in ozone mixing ratios in August 1987. Further laboratory and field studies are needed to quantify better the extent of UV absorption due to nitrated and aromatic aerosols and nitrated aromatic gases.
Article
Full-text available
In situ measurements of atmospheric aerosols have been made from an airborne platform over the remote southern hemisphere ocean in November and December 1995 as part of the First Aerosol Characterization Experiment (ACE 1). A subset of these measurements have been evaluated during three of the flights to characterize the aerosol microphysical and optical properties in the cloud-free, marine boundary. The relationship between the microphysical and optical characteristics in the size range from 0.3 to 20 mum and relative humidity was evaluated. A new technique is introduced by which the scattering coefficient is derived directly from the optical particle counter measurements. The results of this study indicate that changes in particle volume, effective radius, and optical scattering are strongly related to changes in relative humidity (RH). The observations are in very good agreement with laboratory studies of particle volume changes as a function of relative humidity.
Article
Full-text available
To adequately assess the effects of atmospheric aerosols on climate, their optical constants (scattering and absorption coefficients) must be known. The absorption and scattering coefficients of the aerosols are derived from the real and imaginary parts of the complex refractive index and are dependent on their size and chemical composition. Because aerosol properties vary significantly with location, it is difficult to assign values for the absorption and scattering of solar radiation by aerosols in models of global climate change. This study reports a new method of collecting size-fractionated atmospheric aerosol samples for the purpose of directly measuring their transmission and reflectance spectra followed by the determination of the complex refractive index across the entire atmospherically relevant spectral range. The samples were collected with a modified Sierra high-volume cascade impactor with the usual filter collection surfaces replaced with Teflon sheets machined to hold quartz (ultraviolet [UV]/visible transparent) and/or silver chloride (infrared transparent) sample collection plates. Reflectance and transmission spectra can be obtained on the aerosol samples directly as a function of wavelength, from 280 nm to 2.5 m, with an integrating sphere coupled to an UV/visible or a Fourier transform infrared (FTIR) spectrophotometer. The effective real and imaginary components of the refractive index of the bulk sample material can then be approximated, as a function of wavelength, from the sample spectra. Preliminary results are presented for carbon soot samples generated in the laboratory and for standard diesel soot samples in the UV/visible spectral range. These are compared to results obtained for size-fractionated atmospheric aerosol samples collected near Pasco, WA, West Mesa, AZ, and Argonne, IL.
Article
Full-text available
Diesel exhaust has been classified a probable human carcinogen, and the National Institute for Occupational Safety and Health (NIOSH) has recommended that employers reduce workers' exposures. Because diesel exhaust is a chemically complex mixture containing thousands of compounds, some measure of exposure must be selected. Previously used methods involving gravimetry or analysis of the soluble organic fraction of diesel soot lack adequate sensitivity and selectivity for low-level determination of particulate diesel exhaust; a new analytical approach was therefore needed. In this paper, results of investigation of a thermal-optical technique for analysis of the carbonaceous fraction of particulate diesel exhaust are reported. With this technique, speciation of organic and elemental carbon is accomplished through temperature and atmosphere control, and by an optical feature that corrects for pyrolytically generated carbon, or “char,” which is formed during the analysis of some materials. The thermal-optical method was selected because the instrument has desirable design features not present in other carbon analyzers. Although various carbon types are determined, elemental carbon is the superior marker of diesel particulate matter because elemental carbon constitutes a large fraction of the particulate mass, it can be quantified at low levels, and its only significant source in most workplaces is the diesel engine. Exposure-related issues and results of investigation of various sampling methods for particulate diesel exhaust also are discussed. †Disclaimer: Mention of company name or product does not constitute endorsement by the Centers for Disease Control and Prevention. The views expressed in this paper are those of the authors, and do not necessarily reflect the views or policies of the National Institute for Occupational Safety and Health.
Article
Aerosol optical properties in the southeastern United States were measured at two research sites in close horizontal proximity but at different altitudes at Black Mountain (35.66 °N, 82.38 °W, 951 m msl) and Mount Gibbes (35.78 °N, 82.29 °W, 2006 m msl) to estimate the direct radiative forcing in the lowest 1 km layer of the troposphere during the summer of 1998. Measurements of light scattering and light absorption at ambient relative humidity (RH) are categorized by air mass type (polluted continental, marine with some continental influence, continental) according to 48-hour back-trajectory analysis. At a wavelength of 530 nm the average total scattering coefficient (σsp) measured at the valley site was 1.46×10-4 m-1 for polluted continental air masses, 7.25×10-5 m-1 for marine air masses, and 8.36×10-5 m-1 for continental air masses. The ratio of σsp at the mountain site to σsp at the valley site was 0.64, 0.58, and 0.45 for polluted continental, marine, and continental air masses, respectively. The hygroscopic growth factor (σsp(RH = 80%)/σsp(RH = 30%)) was calculated to be almost a constant value of 1.60±0.01 for polluted continental, marine, and continental air masses. As the RH increased from 30% to 80%, the backscatter fraction decreased by 23%. On the basis of these measurements, direct radiative climate forcing (ΔFR) by aerosols in the lowest 1 km layer of the troposphere was estimated. The patterns of ΔFR for various values of RH were similar for the three air masses, but the magnitudes of ΔFR(RH) were larger for polluted continental air masses than for marine and continental air masses by a factor of about 2 due to higher sulfate concentration in polluted continental air masses. The average value of ΔFR(RH = 80%)/ΔFR(RH = 30%) was calculated to be almost a constant value of 1.45±0.01 for all three types of air masses. This implies little dependence of the forcing ratio on the air mass type. The averaged ΔFR for all the observed ambient RHs, in the lowest 1 km layer during the 3-month summer period, was -2.95 W m-2 (the negative forcing of -3.24 W m-2 by aerosol scattering plus the positive forcing of +0.30 W m-2 by aerosol absorption) for polluted continental air masses, -1.43 W m-2 (-1.55 plus +0.12) for marine air masses, and -1.50 W m-2 (-1.63 plus +0.14) for continental air masses. The ΔFR for polluted continental air masses was approximately twice that of marine and continental air masses. These forcing estimates are calculated from continuous in situ measurements of scattering and absorption by aerosols without assumptions for Mie calculations and global mean column burden of sulfates and black carbon (in g m-2) used in most of the model computations.
Article
The wavelength dependence of light absorption by aerosols collected on filters is investigated throughout the near-ultraviolet to near-infrared spectral region. Measurements were made using an optical transmission method. Aerosols produced by biomass combustion, including wood and savanna burning, and by motor vehicles, including diesel trucks, are included in the analysis. These aerosol types were distinguished by different wavelength (λ) dependences in light absorption. Light absorption by the motor vehicle aerosols exhibited relatively weak wavelength dependence; absorption varied approximately as λ-1, indicating that black carbon (BC) was the dominant absorbing aerosol component. By contrast, the biomass smoke aerosols had much stronger wavelength dependence, approximately λ-2. The stronger spectral dependence was the result of enhanced light absorption at wavelengths shorter than 600 nm and was largely reduced when much of the sample organic carbon (OC) was extracted by dissolution in acetone. This indicates that OC in addition to BC in the biomass smoke aerosols contributed significantly to measured light absorption in the ultraviolet and visible spectral regions and that OC in biomass burning aerosols may appreciably absorb solar radiation. Estimated absorption efficiencies and imaginary refractive indices are presented for the OC extracted from biomass burning samples and the BC in motor vehicle-dominated aerosol samples. The uncertainty of these constants is discussed. Overall, results of this investigation show that low-temperature, incomplete combustion processes, including biomass burning, can produce light-absorbing aerosols that exhibit much stronger spectral dependence than high-temperature combustion processes, such as diesel combustion.
Article
We describe a technique to model the radiative properties of mineral aerosols which accounts for their composition. We compile a data set of refractive indices of major minerals and employ it, along with data on mineralogical composition of dust from various locations, to calculate spectral optical and radiative properties of mineral aerosol mixtures. Such radiative properties are needed for climate modeling and remote sensing applications. We consider external mixtures of individual minerals, as well as mixtures of aggregates. We demonstrate that an external mixture of individual minerals must contain unrealistically high amounts of hematite to have a single scattering albedo lower than 0.9 at 500 nm wavelength. In contrast, aggregation of hematite with quartz or clays can strongly enhance absorption by dust at solar wavelengths. We also simulate the daily mean net (solar+infrared) forcing by dust of varying compositions. We found that, for a given composition and under similar atmospheric conditions, a mixture of aggregates can cause the positive radiative forcing while a mixture of individual minerals gives the negative forcing.
Article
Methods for reducing and quantifying the uncertainties in aerosol optical properties measured with the TSI 3563 integrating nephelometer are presented. For nearly all applications, the recommended calibration gases are air and CO2. By routinely characterizing the instrumental response to these gases, a diagnostic record of instrument performance can be created. This record can be used to improve measurement accuracy and quantify uncertainties due to instrumental noise and calibration drift. When measuring scattering by particles, size segregation upstream of the nephelometer at about 1 μm aerodynamic diameter greatly increases the information content of the data for two reasons: one stemming from the independence of coarse and fine particles in the atmosphere, and the second stemming from the size dependence of the nephelometer response. For many applications (e.g., extinction budget studies) it is important to correct nephelometer data for the effects of angular nonidealities. Correction factors appropriate to a broad range of sampling conditions are given herein and are shown to be constrained by the wavelength dependence of light scattering, as measured by the nephelometer. Finally, the nephelometer measurement is nondestructive, such that the sampled aerosol can be further analyzed downstream. Data from two nephelometers operated in series are used to evaluate this procedure. A small loss of super-μm particles (5–10%) is found, while the sub-μm data demonstrates measurement reproducibility within ± 1%.