You are on page 1of 14

Characterization of cooking effluent from seven commercial

kitchen appliances and representative food products.


INTRODUCTION
Characterization of emissions from cooking processes continues to be of interest, with an emphasis
on grease accumulation on surfaces in kitchens and within exhaust systems and its impact on
cleaning and fire potential. Indoor airborne emissions may create potential health hazards to cook
staff and others within the building; widespread dissemination to outdoors will impact atmospheric
chemistry and air quality.
Several regulations on emissions from commercial kitchens have been adopted in the U.S. The
South Coast Air Quality Management District (SCAQMD 1997) implemented Rule 1138 that requires
emission controls on certain chain-driven charbroilers within the district. More recently, the Bay
Area Air Quality Management District (BAAQMD 2007) adopted Rule 2 of Regulation 6 that also
requires emission controls on chain-driven charbroilers beginning January 1, 2009. In addition,
control of underfired charbroiler emissions is scheduled to be phased in between January 1, 2010,
and January 1, 2013.
An annotated bibliography was written (Gerstler et al. 1996) that includes 74 references on
measured cooking effluent and technologies related to its sampling and analysis. Since the
review's publication, several additional studies were reported that add to the archival literature.
Residential applications have been the focus of some research. A series of papers document a study
conducted by the National Exposure Research Laboratory in Reston, Virginia. The first paper
(Wallace 2000) documented several sources of particles and evaluated the levels of polycyclic
aromatic hydrocarbons (PAH) and carbon monoxide in a townhouse with no external ventilation
from the kitchen. Cooking was found to be the major source of particles in the building. The levels
of PAH were deemed insignificant, so they were not monitored in future work at the site. A
subsequent paper (Wallace et al. 2004) focused on particles from 10 nm to 2.5 [micro]m in size.
Forty-four cooking episodes were selected for analysis, as it was determined that cooking (mostly
frying in a skillet) was capable of producing more than ten times the ultrafine particle number
emissions than other activities. Typical cooking episodes produced particles at a rate of
about [10.sup.14] particles/h during a 5-15 min cooking cycle. Ultrafine (<100 nm) and
accumulation mode (0.1-1 [micro]m) particle sources were analyzed for various activities. The
highest mean number concentrations were produced by what was termed "complex cooking" on a
gas stove that generated between 35,000 and 50,000 [cm.sup.-3] compared to less than 1% of this
when no indoor sources were observed. A related study was performed using personal exposure
monitors on 37 participants in their own homes. The highest mean source strengths were identified
from burned food (470 mg/min), grilling (173 mg/min), and using a fry pan (60 mg/min). A group at
the University of Aberdeen Medical School in Aberdeen, Scotland (Dennekamp et al. 2001),
performed a study in which emissions from both electric and gas-operated cooking appliances were
conducted in a small (70 [m.sup.3]) unventilated chamber that simulated a small residence.
Particles from 10 to 500 nm NO and [NO.sub.2] were analyzed. Conclusions drawn from this study
were that potentially toxic concentrations of small particles and very high concentrations of oxides
of nitrogen can be generated without adequate exhaust ventilation.
Effluents from commercial cooking processes were also investigated.Particulate (Kleeman et al.
1999) and organic compounds (Schauer et al.1999) emitted from meat charbroiling were measured

at a commercialkitchen field site. Samples were obtained in the exhaust duct downstreamof the
grease filters installed in the exhaust hood. Frozen and thawedhamburger patties (20% fat) were
cooked on a natural gas-firedcharbroiler. Total mass emission rates were determined to
beapproximately 30 g/kg meat cooked. Peak particle mass concentration wasdetermined using a
micro-orifice uniform deposit impactor (MOUDI) andwas found to be near 0.2 [micro]m. Although
inorganic composition wasinvestigated, the vast majority of the effluent consisted of
organiccompounds. Several nonmethane volatile organic compounds andsemi-volatile organic
compounds were identified including n-alkanoicacids, n-alkenoic acids, and carbonyls.
Personal exposure of the cook staff in 19 commercial kitchens was made in Norway (Svendsen et al.
2002) using glass fiber filter cassettes and a sampling device for aldehydes. Results showed that in
some types of kitchens the concentration of fat aerosols can reach 6.6 mg/[m.sup.3], and the sum of
aldehydes can attain 185 [micro]g/[m.sup.3]. Exposures were found to be greater in small
kitchens than in chain restaurants or hotel kitchens.
A more recent series of measurements conducted at the University of California Riverside CE-CERT
facility was reported (McDonald et al. 2003). Emissions from several appliances were measured, all
in the exhaust duct, after which they were sent through a dilution tunnel and then a residence
chamber to approximate the conditions expected after release to the ambient. Particulate mass
emission factors were considerably less than measured in earlier studies and ranged from 4.5 g/kg
for 21% fat hamburger cooked on a chain-driven char-broiler to 15 g/kg for 25% fat hamburger
cooked on an under-fired charbroiler. The majority of the mass was determined to be organic
carbon with significant amounts of PAHs (primarily naphthalene), lactones, and cholesterol.
Cooking exhaust from 15 commercial kitchens was sampled and analyzed for 13 carbonyl
compounds in Hong Kong (Ho et al. 2006). Various restaurant types were included in the survey
that represented different cooking processes and food styles. The authors concluded that, on a total
mass emissions basis, the top four carbonyls (formaldehyde, acetaldehyde, acrolein and nonanal)
contribute 72% of the carbonyl emissions from commercial kitchens in Hong Kong.
The main goal of this investigation is to extend the data onparticulate and condensable vapor
emissions from commonly usedcommercial kitchen appliances and food products that were
previouslydocumented in the ASHRAE RP-745 final report dated February 9, 1999(Gerstler et al.
1999a). As with the previous study, appropriate foodproducts were selected for each appliance to
provide significant greaseemissions and to be in accordance with corresponding ASTM
Internationaltest protocol requirements. Table 1 provides the seven appliances thatwere tested and
the food product used for each.
The main particle sampling instrument, the personal cascade impactor, and the same grease vapor
sampling instrument, the EPA Method 5 sampling train, were used in this study so the results could
be compared directly with the results obtained in ASHRAE RP-745. Data were obtained both in the
plume from each appliance, as in the earlier study, and also in the exhaust duct with no grease
filters installed in the hood.
One of the observations from the earlier study was that a significant fraction of the particulate mass
emissions can occur in the submicron size range. This was especially true for broilers. In
the previous study, these small particles were captured primarily by the after filter in the personal
cascade impactor and thus their size distribution was not determined. Aerosol measurement
instrumentation has evolved since the earlier study, and now several instruments are available to
measure particles down to a few nanometers in size. A driving factor in the characterization of
ultrafine particle emissions is their impact on human health, as documented in several recent

studies (Oberdorster et al. 2005, 2007). Because of the advancements in instrumentation capability
and the perception that human health issues are associated with ultrafine particles, two scanning
mobility particle sizers (SMPS) were added to the instrumentation package that cover the range
from 20 nm to 0.8 [micro]m. Thus, particles from 20 nm to 15 [micro]m in diameter are captured
and quantified.
Table 1. List of Appliances Tested and Corresponding Food Products Type Brand And Model Food
Product Conveyor Broiler Nieco model 980 Frozen 1/8 lb hamburger patties 10% fat, 55%
moisture Clamshell Griddle Garland model Frozen 1/4 lb MWG-9501 hamburger patties 15% fat,
48% moisture Conveyor Pizza Oven Middleby Marshall Frozen 17.75 oz 12" model PS360-WB
diameter thin crust pepperoni pizzas Overfired Broiler Vulcan model 5 oz sirloin steaks, Sunglow
IR-71P 2.0% fat, 72.3% moisture Electric Steamer Stellar model 5 oz boneless Steam Altair II
skinless chicken breasts Mesquite Solid Fuel Holstein Frozen 1/4 lb Broiler Manufacturing
hamburger patties 10% model Charcoal fat, 55% moisture Country Club Custom 36" length Gas
Chinese Wok Jade model JCR-1 5 oz boneless skinless chicken breasts diced into 1" cubes with
peanut oil
TEST FACILITY
The test kitchen is located within the Mechanical Engineering Building at the University of
Minnesota and was constructed and used in the previous ASHRAE RP-745 (Gerstler et al. 1999a)
study. A schematic drawing of the facility and associated instrumentation is provided in Figure 1
and a list of instrumentation is provided in Appendix A. The construction is of steel frame with floor
dimensions 10 x 10 ft (3.05 x 3.05 m) with an inside height of 9 ft (2.74 m). Wood 2 x 4 studs
are attached to the steel frame and support 5/8 in. (1.59 cm) fire-rated wallboard. Twenty-gauge
stainless steel panels cover the entire back wall behind the appliances. The floor is covered with a
single sheet of linoleum cut to fit the interior dimensions of the space. The wall opposite the
appliance location is constructed of a two-part removable wood frame covered by window screen.
The frames are attached to the chamber walls by hinges with removable pins for moving the
appliances in and out of the facility. The screen allows makeup air to be taken from the remaining
laboratory space at low velocity so as not to interfere with the thermal plume above the appliance or
performance of the exhaust hood.
[FIGURE 1 OMITTED]
An 8 ft (2.44 m) long by 4 ft (1.22 m) wide type 1 wall-mounted canopy listed ventilation hood is
bolted to the steel frame that supports the ceiling of the chamber. The hood has an internal depth of
2 ft (0.61 m) and is placed against the back wall of the chamber with the opening at a height of 6.5
ft (1.98 m) above the floor. A 16 in. (40.64 cm) diameter collar connects the hood to a horizontal
round stainless steel exhaust duct of the same diameter. The duct runs horizontally to a 16 in.
(40.64 cm) 1 hp centrifugal exhaust fan modified by the manufacturer to operate in a horizontal
position. The fan exhausts to the outside air. Tempered makeup air is provided by a separate airhandling unit installed in the same room and by drawing additional air from the large building.
Appliance connections along the back wall include one three-phase 208 V electrical outlet, a 4-plug
115 V grounded outlet, and a 1 in. (2.54 cm) natural gas line. In addition, a 4-plug 115 V grounded
outlet is mounted inside the right hand wall for instrumentation power as is the variable speed
drive for the exhaust fan. A fire extinguisher and a personnel access door are on the left wall near
the removable screen panel.
Each appliance was located at the center of the exhaust hood and at least 6 in. (15.2 cm) back from

the front of the hood for appropriate capture and containment of the effluent plume. Whenever
possible, trained personnel who were familiar with operation of each appliance were present to
supervise the connections and to ensure the appliance was calibrated and operating normally before
any cooking was initiated. Figure 2 shows the solid-fuel broiler positioned under the exhaust
hood before the EPA Method 5 sampling train and the attached personal cascade impactors (PCIs)
were moved into position.
[FIGURE 2 OMITTED]
INSTRUMENTATION
Particles were captured and classified in the size range of 0.5 to 15 [micro]m using Marple model
298 PCIs with model 290 IA in-line adaptors. Short sections of copper tubing, 0.183 in. (4.65 mm)
I.D., were given a tapered inlet and a 2.48 in. (63 mm) bend radius to form the 90[degrees]
sampling inlets. Substrates were 1.34 in. (34 mm) mylar. Final filters were either 1.34 in. (34 mm)
PVC membrane with 5 [micro]m pore size or 1.34 in. (34 mm) glass fiber filters. The air drawn
through the impactors was then sent either to the EPA Method 5 sampling train to remove the
condensable grease or to a vacuum pump.
Particles between 20 nm and 0.8 [micro]m in size were analyzed by two SMPS, one sampling in the
plume and one in the exhaust duct. Sampling inlets were fabricated from copper tubing similar to
those for the impactors. Approximately 6 in. (15 cm) from the inlet, the sample was diluted by a
factor of 10 with filtered dry air. The diluted sample was then sent to the SMPS for analysis of the
particle size distribution. Each scan over the particle size range required approximately four
minutes. For appliances with variable emissions, the scanning results must be averaged over
several cooking cycles to provide data comparable to appliances with more steady emissions.
A third sample of the effluent in the plume was taken through 1.85 in. (47 mm) open-face quartz
filters for subsequent chemical composition analysis. A second filter was placed downstream of the
primary filter to capture anything that passed through it. Aerosol samples in the exhaust duct were
captured in a separate copper sampling line and sent to an aerosol time-of-flight mass spectrometer
(ATOFMS) for determination of particle chemistry by size. The chemistry results will be provided in
a separate paper.
A sketch of the instrumentation layout in the test facility is shown in Figure 1. All the
instrumentation associated with sampling in the plume was located within the test kitchen. The
instrumentation used to sample and characterize effluent in the exhaust duct was located in the
surrounding laboratory space. The dashed lines in Figure 1 indicate the instruments that were used
for only some of the tests. A photograph of a representative effluent sampling setup for the
conveyor broiler is shown in Figure 3. All three sampling inlets (PCI, SMPS, filter holder) were
located as close to the center of the effluent plume as possible.
[FIGURE 3 OMITTED]
Other instrumentation included type T thermocouples for air temperature measurement, type K
thermocouples for appliance temperature measurements, a portable hot film anemometer for air
velocity measurements in the exhaust duct, an optical particle counter for checking the uniformity
of aerosol distribution in the exhaust duct, and an electrical power data logger and a natural gas
flowmeter to measure the energy input rates into the appliances.
PROCEDURE

After each appliance was installed and operating correctly, several preliminary cooking runs were
made following the appropriate ASTM procedure developed for that appliance. The results from
these preliminary tests were used chef hire to determine the amount of food product to use, the
cooking time to achieve the desired weight loss and/or internal temperature, and the number of
batches or length of cooking needed to capture sufficient effluent for analysis.
Each of the impactor substrates was desiccated and weighed prior to installation into the substrate
holders and then into the impactor assembly. Clean after filters were installed and the impactors
connected to the end of the Method 5 heated probe or the vacuum pump. The completed Method 5
assembly with the impactor at the front end was then positioned so that the inlet was as close to the
center of the effluent plume as possible.
The center of the plume was determined by mapping the air temperature near the bottom of the
exhaust hood and determining the point with the highest temperature. Clean quartz filters were
installed into the filter holders, and the filter assemblies were positioned under the hood and
connected to the vacuum pump. An SMPS, dilution air system, and computer were all located inside
the test kitchen opposite the personnel door. The sampling inlet was repositioned slightly
from appliance to appliance as the plume center changed location.
Sampling from the exhaust duct was accomplished using an isokinetic probe positioned in the
centerline of the duct. A second PCI assembly was connected to a flow-calibrated laminar flowmeter
to determine the correct sampling airflow rate. This was connected to a vacuum pump except the
runs where the method 5 sampling train was used to determine grease vapor concentration in the
exhaust duct. The SMPS shared the same sampling line with the PCI but no dilution air was used as
the concentration in the exhaust duct was sufficiently low so that dilution was not necessary.
Once all the instrumentation was positioned, the appliance was turned on and allowed to come to
operating temperature. At the beginning of the cooking, all vacuum pumps were turned on
simultaneously so that all of the instruments sampled during the same time interval. The
SMPS instruments were restarted several times during a run to take sweeps over the particle size
range. Once sufficient time had elapsed, the cooking was terminated and all pumps shut off. The
impactors were then disassembled and the stage substrates allowed to dry in a desiccator for at
least 24 hours. Once the substrates were dry, the final weight measurements were made that were
used to determine the mass of particles collected for the size range that corresponded to each
impactor stage. The Method 5 impingers were washed out with acetone and placed
into preweighed beakers. The solvent was allowed to evaporate, and the final weight was measured
to determine the amount of condensable vapor collected. The quartz filters were removed from the
filter holders, put into sealed bags, and placed in a freezer for subsequent chemical analysis. The
data from the SMPS measurements were curve fitted using a commercial software package that
assumed a log-normal distribution and determined the mean particle size and geometric standard
deviation.
A minimum of three tests were run as described above for each appliance, with the Method 5
sampling train used to determine the condensable grease effluent in the plume under the hood. At
least one run was then made with the Method 5 used to sample effluent in the exhaust duct.
RESULTS AND DISCUSSION
Results are separated into total grease mass particulate and vapor levels measured in the plume
and exhaust duct and particle size distributions measured by both the personal cascade impactor
and the scanning mobility particle sizer in the plume and duct.

Fraction of the Plume that was Sampled


In the ASHRAE RP-745 (Gerstler et al. 1999a) study, small particles collected on the lower stages of
an electrical low pressure impactor (ELPI) or MOUDI in the plume and in the exhaust duct were
used to determine the dilution ratio. This is needed to determine the fraction of the plume that was
sampled and to calculate the total mass emissions in the plume. However, for this study there were
three possible methods for determining the dilution ratio: (1) SMPS integrated
number concentrations in the plume and duct, (2) particle mass collected on the PCI after filters in
the plume and duct, or (3) mass of condensable grease vapor collected using the EPA Method 5
apparatus in the plume and duct. The assumption is made that negligible transfer of mass
from particle-to-vapor phase and vice versa occurs between the plume sampling point and the
sampling point in the exhaust duct. This is a reasonable approximation, as the effluent time interval
between the two sampling locations is approximately one second, and the effluent is already diluted
and cooled by the time it reaches the sampling point in the plume. A similar observation was made
by Schauer (Schauer et al. 1999), where the transfer of semivolatile organic compounds from meat
cooking appeared to be prohibited from transferring between the gas phase and particle phase. The
authors attributed this to the solid character of meat fat at room temperature, which has a relatively
low vapor pressure. When looking at each of these possibilities, it was found that the SMPS raw
number concentrations were sometimes greater in the duct than in the plume. This occurred when
the plume was not well defined-for example with the clamshell griddle, where the center of the
plume varied from side to side, and with the conveyor pizza oven, where the effluent was emitted
out of both ends of the oven. In these cases the results led to dilution ratios less than one and
indicated that normalized mass concentrations in the duct would be greater than in the plume,
which is unrealistic. The second possibility was to use the mass collected on the PCI after filters.
Here it was found that, in some cases, the dilution ratios were large enough that calculated
normalized condensable grease concentrations in the duct would be greater than in the plume,
which is also unrealistic. The third possibility of using dilution ratios calculated from the
condensable grease vapor collected in the plume and duct provided the most realistic results, where
the normalized particle concentrations in the plume were always greater than or equal to those in
the duct. Determining dilution ratios from collected grease vapor in the plume and duct seems to be
reasonable since there is no significant grease vapor removal mechanism between the two sampling
locations. Therefore, the dilution ratio for each appliance was calculated by dividing the
concentration of condensable grease vapor collected in the plume by the concentration of
condensable grease vapor collected in the duct.
Grease Particulate and Vapor Levels in Plume and Exhaust Duct
The total normalized grease mass emissions measured in the plume and exhaust duct from each
appliance are shown in Figures 4 and 5, respectively, with the results broken down into vapor,
PM2.5, PM10-PM2.5 and particles larger than 10 [micro]m. Numerical values are provided
in Appendix B--Table B1 for the results in the plume and Table B2 for the duct. Note that no grease
filters were used in the hood during these tests, so the emissions in the exhaust duct are different
than what would be expected in a real cooking scenario where grease filters or extractors are
present. This was a requirement of the initial RFP and provided information on natural effluent
elimination by the exhaust system between the plume and exhaust duct without the use of
grease filters. Results from the gas conveyor broiler cooking hamburger have an emission
distribution that is comparable to the previous results for the gas and electric underfired broilers
cooking hamburger (Kuehn et al. 1999). The total grease mass emissions from the clamshell
griddle, conveyor pizza oven, and overfired broiler cooking steak are relatively low and comparable
to several of the lower-emitting appliances tested previously. Grease vapor is the predominant
emission from all these appliances. No reliable PCI or SMPS data were obtained in the plume of the

steamer, as the high water content precluded accurate measurements there. The emissions from the
solid fuel broiler cooking hamburger are similar in mass distribution but more than twice the
quantity of those from the gas underfired broiler cooking hamburger from the RP-745 study. The
wok generated huge amounts of large particles in the plume primarily caused by spatter or
mechanical particle generation that caused the total emission to be a factor or two larger than the
total from any other appliance.
[FIGURE 4 OMITTED]
[FIGURE 5 OMITTED]
The total particulate mass emissions in the plume for three of the appliances--the clamshell griddle,
conveyor pizza oven, and overfired broiler--are very similar to those measured in the exhaust duct.
The particulate mass emissions for the conveyor broiler and the solid-fuel broiler are significantly
higher--twice as high in the plume as in the exhaust duct. This can be explained by the higher
proportion of mass in the largest particle size measured that indicates that some of these particles
do not make it to the exhaust duct sampling probe and may be lost by impaction or settling. The
wok has much higher particulate mass emission in the plume than in the exhaust duct by a factor of
about five. Again, this can be explained by the overwhelming amount of large grease particles in
the plume that are not found in the exhaust duct.
McDonald et al. (2003) compared their PM 2.5 mass emission results with those of previous
investigators. The results closest to the present study are the emissions using a conveyor broiler
with hamburger. They obtained a value of 4.5 lb (kg)/1000 lb (kg) food product, whereas Norbeck
et al. (1997) measured 7.4 lb (kg)/1000 lb (kg) food product. Both used hamburger with 21% fat.
The present results were obtained with 10% fat hamburger and give a PM 2.5 value of 8.0 lb
(kg)/1000 lb (kg) food product. Both of these previous studies used considerably different methods
than the particle/ vapor sampling procedures used here, as they were focused on emissions into the
environment. McDonald obtained the sample from the exhaust duct and passed it into a dilution
tunnel with a dilution factor of 30:1 followed by a mixing chamber with a 90 s residence time to
simulate the conditions expected after release into the atmosphere. In the present study, samples
were passed as quickly as possible into the particle and vapor sampling instruments to characterize
the nature of the effluent stream in the plume and duct, which does not necessarily correspond to
the ultimate fate in the atmosphere. When the masses associated with other particle sizes from this
study are added (2.8 lb [kg]/1000 lb [kg] food product) and the condensable vapor is added (16.9 lb
[kg]/ 1000 lb [kg] food product), the total normalized mass emissions measured here become 27.6
lb (kg)/1000 lb (kg) food product.
Particle Size Distributions
The particle size distributions in the plume and exhaust duct were determined for each appliance
using the PCI for the larger sizes and the SMPS for the smaller sizes. Particles that impacted inside
the PCI sampling inlet tube were removed by swabbing with acetone, and the results were added to
the mass removed from the impactor inlet and Stage 1. Thus the smallest particles included in this
total were larger than the cut size for Stage 1. However, the largest size is unknown. As in the
previous study, the maximum particle size was assumed to be 100 [micro]m. The results for Stages
2 through 8 are well characterized, as the cut sizes for each of these stages is known as a function
of airflow rate. The largest particles captured on the final filter are equal to the cut size for Stage 8
of the impactor. However, the smallest particle size is unknown. The smallest size is assumed to be
10 nm (0.01 [micro]m), except where noted. The results from the SMPS are used to provide more
detailed information on these small particles. It is also important to note that the impactor provides

mass concentration data, whereas the SMPS provides number concentration data. These results
can be combined, if the particle density is assumed, so that the number concentration results can
be converted to mass concentration for direct comparison. All the particle size results are included
in the project final report (Kuehn et al. 2008). Results from the conveyor broiler are presented here
as representative.
Figure 6 shows the particle size data from three runs with the personal cascade impactors sampling
from the plume and exhaust duct for the conveyor broiler cooking hamburger. The results are given
in terms of particle mass per unit volume. The sampling airflow rate changed slightly from run to
run, so the cut sizes of the impactor stages are slightly different for each of the three runs. The
general size distribution is similar between the plume and exhaust duct, although there are fewer
large particles in the duct. Note that the concentration of the smaller particle sizes decreases
dramatically between the plume and the duct by about a factor of four. This is to be expected
because of the large amount of dilution air brought in by the hood.
[FIGURE 6 OMITTED]
The impactor and its sampling inlet capture particles that range in size from 0.5 to about 100
[micro]m. Particles smaller than 0.5 [micro]m are captured on an after filter that does not provide
resolved particle size information. To resolve the particle size distribution in the size range, SMPSs
were used. Figure 7 shows the corresponding results from the SMPS measurements taken in the
plume and in the exhaust duct. The plots show number concentration mean and standard deviation
for each SMPS channel calculated from a minimum of three scans per run with three runs in the
plume and four in the duct. Therefore, at least nine scans were taken in the plume and 12 in the
duct that were used in calculating the means and standard deviations.
[FIGURE 7 OMITTED]
The size distributions measured with the SMPS instruments show that the maximum particle
number concentrations occur near 100 nm (0.1 [micro]m) in size. This is the result of heterogeneous
nucleation, where vapor condenses onto small nuclei, and these droplets continue to grow until
they reach about 100 nm in size. Similar size distributions have been measured in many previous
aerosol studies. This physical phenomenon is described in more detail in the text by Hinds (Hinds
1999).
To offer additional information about the SMPS data, a commercial data reduction program was
used to fit a curve to the mean values under the assumption that the data are log-normally
distributed. The results of the fit indicate that the geometric mean diameters are 0.138 and 0.144
[micro]m, and the geometric standard deviations are 1.73 and 1.65 for the average results in the
plume and the duct, respectively. This indicates that the size distributions are nearly identical
between the plume and duct sampling points and corroborates the assumption made earlier that
negligible grease transfer between particle and vapor phases occurs between the two locations.
However, the concentrations in the duct are reduced significantly because of dilution. Geometric
mean particle diameters and corresponding geometric standard deviations from the SMPS data for
all appliances tested are provided in Table 2. The conveyor broiler and the solid fuel broiler have
mean diameters of approximately 0.15 [micro]m, whereas all the others are less than
100 [micro]m.
The particle mass concentration results obtained with the PCI shown in Figure 6 and the small
particle number concentration data obtained using the SMPS shown in Figure 7 can be compared if
the particles are given an assumed density so that the particle number concentration data can be

converted into mass concentration. Using an assumed grease particle density of 0.90 g/c[m.sup.3],
the SMPS results are overlayed onto the PCI results for the plume and exhaust duct in Figure 8.
[FIGURE 8 OMITTED]
Table 2 Particle Mean Diameters and Geometric Standard Deviations from the Scanning Mobility
Particle Sizer Data from all the Appliances Tested Plume Exhaust Duct Geometric
Geometric Appliance Mean, Standard Mean, Standard [micro]m Deviation [micro]m
Deviation Conveyor Broiler 0.138 1.73 0.144 1.65 Clamshell Griddle 0.086 2.14 0.098
1.76 Conveyor Pizza Oven 0.036 1.53 0.038 1.53 Overfired Broiler 0.044 1.57 0.049 1.54 Steamer
<0.02 Solid Fuel Broiler 0.173 1.79 0.145 1.66 Wok 0.054 2.14 0.055 1.88
A lower limit of 0.01 [micro]m (10 nm) has been assumed for particles captured by the final filter
used on the impactor, as was assumed in the previous ASHRAE study (Gerstler et al. 1999b).
However the results from the scanning mobility particle sizer now provide much better information
on the small particle size range. Changing the assumed lower particle size limit from 0.01 to 0.15
[micro]m for the final filter used with the conveyor broiler and more consistent with the results
from the SMPS and replotting the results in the plume given in Figure 8a provides the results
shown in Figure 9. The mass concentration results provided by combining the data from the two
instruments are now more consistent. Similar adjustments can be made for the other appliances
that were tested. However, two assumptions need to be made: (1) a grease particle density and (2) a
lower particle size limit for the final filter in the personal cascade impactor.
[FIGURE 9 OMITTED]
SUMMARY AND CONCLUSIONS
The total grease mass emissions in the plume were found to be comparable to the emissions
documented in the ASHRAE RP-745 (Gerstler et al. 1999a) study for some of the appliances. Results
from the conveyor broiler agree well with previous results from the underfired broiler cooking
hamburger. Emissions from the conveyor pizza oven were the lowest at approximately 2.5 lb
(kg)/1000 lb (kg) food product and agreed with the results from the previous pizza ovens. Emissions
from the clamshell griddle were similar to those from the electric griddle tested previously. The
overfired broiler cooking steak generated between 10 and 15 lb (kg)/ 1000 lb (kg) food cooked that
consisted primarily of grease vapor and compared well with the results from the underfired
broilers cooking chicken in the earlier study.
Both the solid-fuel broiler cooking hamburger and the wok cooking diced chicken breast in peanut
oil generated huge amounts of large particles in the plume. However very little of this was found in
the exhaust duct, although no filters were installed in the hood.
Although large amounts of grease mass corresponding to particles larger than 10 [micro]m were
found in the plume from the solid fuel broiler and the wok, nearly all of the grease mass emission in
the exhaust duct was found to be in the vapor phase or associated with particles smaller than 1
[micro]m in size. Neither of these effluent components can be easily removed by inertial impaction
utilized by most current technology grease filters. Thus, novel grease removal technologies should
be developed and implemented to better control grease emissions from commercial kitchens using
the appliances tested here. High concentrations of ultrafine particles similar to what
were measured in this study have been shown to increase human health risk associated with
respiratory exposure. This issue should be investigated further to determine what levels of emission
control are needed based on a variety of factors, including economic impacts on the food

service industry and the risks to worker health.


ACKNOWLEDGMENTS
We would like to acknowledge several companies and individuals who helped make this research
possible:
* Metal-Fab for donating the exhaust duct
* Greenheck for the use of the optical particle counter, exhaust hood, and fan
* Burger King for supplying the conveyor broiler, a technician to assist in the proper setup, and the
food product used (hamburger patties)
* McDonalds, Garland, Martin Brower, and OSI for supplying the clamshell griddle and the food
product (hamburger patties)
* PG&E Food Service Technology Center for supplying the conveyor pizza oven, overfired broiler,
steamer, solid-fuel broiler and wok
* Kevin Stockman for cooking during the wok tests
* Fisher Nickel, Inc. for coordinating the cofunding needed for the chemical analyses
Finally, we would like to thank the members of ASHRAE TC 5.10 and the members of the RP-1375
Project Monitoring Subcommittee for their suggestions, guidance, and interest in this
investigation.
REFERENCES
ASTM Standard Test Method for Performance of Steam Cookers, F 1484-05
ASTM Standard Test Method for Performance of Double-Sided Griddles, F 1605-95
ASTM Standard Test Method for Performance of Underfired Broilers, F 1695-03
ASTM Standard Test Method for Performance of Conveyor Ovens, F 1817-97
ASTM Standard Test Method for Performance of Chinese (Wok) Ranges, F 1991-99
ASTM Standard Test Method for Performance of Upright Overfired Broilers, F 2237-03
ASTM Standard Test Method for Performance of Conveyor Broilers, F 2239-03
BAAQMD. 2007. Regulation 6: Particulate matter, Rule 2: Commercial cooking equipment. Bay Area
Air Quality Management District, San Francisco, CA.
Dennekamp, M., S. Howarth, C.A.J. Dick, J.W. Cherrie, K. Donaldson, and A. Seaton. 2001. Ultrafine
particles and nitrogen oxides generated by gas and electric cooking. Journal of Occupational and
Environmental Medicine 58:511-16.

Gerstler, W.D., T.H. Kuehn, D.Y.H. Pui, J.W. Ramsey, and M.P. Doerr. 1996. Identification and
characterization of effluents from various cooking appliances and processes as related to optimum
design of kitchen ventilation systems. ASHRAE RP-745, Phase 1 Final Report. Atlanta: American
Society of Heating, Refrigerating and Air Conditioning Engineers, Inc.
Gerstler, W.D., T.H. Kuehn, D.Y.H. Pui, J.W. Ramsey, M. Rosen, R.R. Carlson, and S.D. Petersen.
1999. Identification and characterization of effluents from various cooking appliances and processes
as related to optimum design of kitchen ventilation systems. ASHRAE RP-745, Phase 2 Final
Report. Atlanta: American Society of Heating, Refrigerating and Air Conditioning Engineers, Inc.
Gerstler, W.D., T.H. Kuehn, D.Y.H. Pui, and J.W. Ramsey. 1999. Measurements of the effluent from
hamburger cooked on a gas underfired broiler. ASHRAE Transactions 105(2):303-15.
Hinds, W.C. 1999. Aerosol Technology, 2nd ed. Hoboken, NJ: John Wiley & Sons.
Ho, S.S.H., J.Z. Yu, K.W. Chu, and L.L. Yeung. 2006. Car-bonyl emissions from commercial cooking
sources in Hong Kong. Journal of Air & Waste Management Association 56:1091-98.
Kleeman, M.J., J.J. Schauer, and G.R. Cass. 1999. Size and composition distribution of fine
particulate matter emitted from wood burning, meat charbroiling and cigarettes. Environmental
Scince Technology 33:3516-23.
Kuehn, T.H., W.D. Gerstler, D.Y.H. Pui, and J.W. Ramsey. 1999. Comparison of emissions from
selected commercial kitchen appliances and food products. ASHRAE Transactions 105(2):128-41.
Kuehn, T.H., B.A. Olson, J.W. Ramsey, J. Friell, and J.M. Rocklage. 2004. Development of a standard
method of test for commercial kitchen grease removal systems. Final Report, Department of
Mechanical Engineering, University of Minnesota, Minneapolis, MN.
Kuehn, T.H., B.A. Olson, J.W. Ramsey, J. Rocklage. 2008. Characterization of effluents from
additional cooking appliances. Final Report, ASHRAE 1375-RP, American Society of Heating,
Refrigerating and Air-Conditioning Engineers, Inc., Atlanta.
McDonald, J.D., B. Zielinska, E.M. Fujita, J.C. Sagebiel, J.C. Chow, and J.G. Watson. 2003. Emissions
from charbroiling and grilling of chicken and beef. Journal of Air & Waste Management
Association. 53:185-94.
Oberdorster, G., E. Oberdorster, and J. Oberdorster. 2005. Nanotoxicology: An emerging discipline
evolving from studies of ultrafine particles. Environmental Health Perspectives 113:823-39.
Oberdorster, G., V. Stone, and K. Donaldson. 2007. Toxicity of nanoparticles: A historical
perspective. Nanotoxicol-ogy 1(1):2-25.
Olson, D., and J.M. Burke. 2006. Distributions of PM2.5 source strengths for cooking from the
Research Triangle Park particulate matter panel study. Environmental Science and Technology
40:163-69.
E.H. Pechan & Associates, Inc. 2003. Technical Memorandum: Methods for developing a national
emission inventory for commercial cooking processes. Report prepared for the Emission Factor and
Inventory Group, Emissions Monitoring and Analysis Division, Office of Air Quality Planning and
Standards, U.S. EPA, Research triangle Park, NC.

SCAQMD. 1997. Rule 1138: Control of emissions from restaurant operations. South Coast Air
Quality Management District, Diamond Bar, CA.
Schauer, J.J., M.J. Kleeman, G.R. Cass, and B.R.T. Simonet. 1999. Measurement of emissions from
air pollution sources. 1. C1 through C29 organic compounds from meat charbroiling. Environmental
Science and Technology 33:1566-77.
Svendsen, K., H.N. Jensen, I. Sivertsen, and A.K. Sjaastad. 2002. Exposure to cooking fumes in
restaurant kitchens in Norway. Annual of Occupational Hygiene 46(4):395-400.
Wallace, L. 2000. Real-time monitoring of particles, PAH, and CO in an occupied townhouse.
Applied Occupational and Environmental Hygiene 15(1):39-47.
Wallace, L.A., S.J. Emmerich, and C. Howard-Reed. 2004. Source strengths of ultrafine and fine
particles due to cooking with a gas stove. Environmental Science and Technology 38:2304-11.
Wallace, L. 2006. Indoor sources of ultrafine and accumulation mode particles: size distributions,
size resolved concentrations, and source strengths. Aerosol Science and Technology 40:348-60.
APPENDIX A
Particulates
1. Personal Cascade Impactor (PCI)
a. Marple model 298 with model 290 IA in-line adaptor
b. 90[degrees] sampling probe: copper tubing, machine tapered inlet, 2.48 in. (63 mm) bend radius,
0.183 in. (4.65 mm) inside diameter
c. 1.34 in. (34 mm) mylar substrates: model c-290-MY
d. 1.34 in. (34 mm) PVC membrane after-filters: 5 um pore size, model F-290-P5
e. 1.34 in. (34 mm) glass fiber after-filters: Spectro grade, type A, Gelman filtration media
2. Total Filter Sample
a. 1.85 in. (47 mm) Millipore open-face filter holder
b. 1.85 in (47 mm) quartz filters.
3. Scanning mobility particle sizer (SMPS) used for plume measurements
a. Differential mobility analyzer, TSI model 3071
b. Condensation particle counter, TSI model 3025
4. Scanning mobility particle sizer (SMPS) used for duct measurements
a. Differential mobility analyzer, manufactured in house

b. Condensation particle counter, TSI model 3010


5. Aerosol time-of-flight mass spectrometer (ATOFMS), TSI series 3800
6. Optical particle counter (OPC), Climet model Spectro 0.3 airborne
Condensable Grease Vapor
1. EPA Method 5 sampling train, Graseby Andersen universal stack sampler
2. Condensation vapor monitor constructed in house
Temperature
1. Air: Type T thermocouple, 24 gauge wire, Teflon PFA insulation (500[degrees]F, 260[degrees]C).
2. Appliance: Type K thermocouple, 20 gauge wire, glass braid insulation (900[degrees]F,
482[degrees]C).
3. General digital multimeter, Keithley 132F TRMS Multimeter
Velocity
1. Portable hot-wire anemometer, TSI model 8330 VelociCheck
Appliance Energy Measurement
1. Electrical power data logger, Synergistic model C 180
2. Natural gas flowmeter with 1/8 [ft.aup.3] digital resolution
Analytic
1. Analytic microbalance, Cahn C-31, 0-25 mg, 0.1 mg resolution
2. Fine analytic balance, Sartorius B-120S, 0-110g, 0.1 mg resolution
3. Course analytic balance, Sartorius L-610, 0-610 g, 0.01 g resolution
4. Kitchen scale, Pelouze model Y32R, 0-32 oz resolution
APPENDIX B
Table B1. Normalized Grease Mass Emission Results for Six Appliances Measured in the Plume
Below the Hood lb. Emissions/1000 lb. Food Product Appliance Dp > 10 2.5 um < Dp Dp [less
than Condensable Total [micro]m [less than or or equal to] Vapor equal to] 10 2.5 [micro]m
[micro]m Conveyor 22.1 0.69 10.5 16.9 50.2 Broiler (Hamburger) Clamshell 2.94 0.58 0.70 9.15
13.4 Griddle (Hamburger) Conveyor 0.19 0.02 0.03 2.39 2.64 Oven (Pizza) Overfired 2.01 0.19
1.16 7.44 10.8 Broiler (Steak) Solid-Fuel 55.2 2.68 35.4 48.9
142 Broiler using Mesquite (Hamburger) Chinese Wok 188 4.89 11.2 43.0
247 (Cubed Chicken Breast in Peanut Oil) Table B2. Normalized Grease Mass Emission Results

for All Seven Appliances Measured in the Exhaust Duct with No Grease Filters Present lb.
Emissions/1000 lb. Food Product Appliance Dp > 10 2.5 um < Dp Dp [less than Condensable
Total [micro]m [less than or or equal to] Vapor equal to] 10 2.5 [micro]m [micro]m Conveyor
2.35 0.41 7.97 16.9 27.6 Broiler (Hamburger) Clamshell 0.81 0.41 0.37 9.15
10.7 Griddle (Hamburger) Conveyor 0.17 0.03 0.05 2.39 2.64 Oven (Pizza) Overfired 1.56 0.14
0.78 7.44 9.93 Broiler (Steak) Steamer 1.16 0.07 0.06 13.7 15.0 (Chicken Breast) Solid-Fuel
0.46 1.69 21.4 48.9 72.5 Broiler Using Mesquite (Hamburger) Chinese Wok 4.03 3.00 5.62 43.0
55.6 (Cubed Chicken Breast in Peanut Oil)
Thomas H. Kuehn, PhD, PE
Fellow ASHRAE
James W. Ramsey, PhD
Member ASHRAE
Bernard A. Olson, PhD
Joshua M. Rocklage
Student Member ASHRAE
This paper is based on findings resulting from ASHRAE Research Project RP-1375.
Thomas H. Kuehn is a professor and director of the Environmental Division, Bernard A. Olson is a
research associate, James W. Ramsey is a professor and associate department head, and Joshua M.
Rocklage is a research assistant in the Department of Mechanical Engineering, University of
Minnesota, Minneapolis, MN.
http://www.thefreelibrary.com/Characterization+of+cooking+effluent+from+seven+commercial+ki
tchen...-a0201591022

You might also like