You are on page 1of 43

The S ociety for engineering in agricultural, food, and biological systems

Paper Number: 01-4090 An ASAE Meeting Presentation

Sampling and Measurement of Ammonia Concentration at Animal Facilities A Review


Ji-Qin Ni, Ph.D., Senior Research Associate
Agricultural & Biological Engineering Department, Purdue University, West Lafayette, IN 47907, USA, jiqin@ecn.purdue.edu

Albert J. Heber, Ph.D., P.E., Associate Professor


Agricultural & Biological Engineering Department, Purdue University, West Lafayette, IN 47907, USA, heber@purdue.edu

Written for presentation at the 2001 ASAE Annual International Meeting Sponsored by ASAE Sacramento Convention Center Sacramento, California, USA July 30 August 1, 2001
Abstract. Scientific understanding and technical control of ammonia (NH3) at animal facilities largely depend on sampling and measurement techniques. Good sampling/measurement techniques provide high quality data that are essential in the study and abatement of NH3 emission. For many years until now, there have been only a few publications that address the application of these techniques at animal facilities under field conditions. This review emphasizes the sampling and measurement of NH3 applied under field conditions and draws an updated picture of the state-of-the-art of NH3 concentration measurement at animal facilities. Four types of sampling techniques and nine groups of measuring devices for NH3 concentration are described and summarized. Variables of measurement, and principles, procedures, advantages and disadvantages of various sampling/measurement techniques are reviewed. Recommendations are made for future research including development of methodologies and standards. Keywords. Air quality, Environment, Agriculture, Instrument, Sensor

The authors are solely responsible for the content of this technical presentation. The technical presentation does not necessarily reflect the official position of the American Society of Agricultural Engineers (ASAE), and its printing and distribution does not constitute an endorsement of views which may be expressed. Technical presentations are not subject to the formal peer review process by ASAE editorial committees; therefore, they are not to be presented as refereed publications. Citation of this work should state that it is from an ASAE meeting paper. EXAMPLE: Authors Last Name, Initials. Title of Presentation. ASAE Meeting Paper No. xx-xxxx. St. Joseph, Mich.: ASAE. For information about securing permission to reprint or reproduce a technical presentation, please contact ASAE at hq@asae.org or 616.429.0300 (2950 Niles Road, St. Joseph, MI 49085-9659 USA).

Table of Contents
ABBREVIATIONS AND ACRONYMS..................................................................................................II INTRODUCTION....................................................................................................................................... 1 A GENERAL VIEW OF AMMONIA MEASUREMENT...................................................................... 2 AMMONIA SAMPLING ........................................................................................................................... 4 LOCATION AND TIME OF SAMPLING ......................................................................................................... 4 METHODS AND DEVICES OF SAMPLING .................................................................................................... 5 Closed Sampling: Sampling Chambers................................................................................................. 5 Point Sampling and Time Control: Air Stream Controller ................................................................... 7 Open-path Sampling: Optical Detection Device................................................................................... 7 Sampling Air Volume Control............................................................................................................... 8 AMMONIA CONCENTRATION MEASURING DEVICES ................................................................ 8 CLASSIFICATION........................................................................................................................................ 8 Wet and Dry .......................................................................................................................................... 9 Sensitivity .............................................................................................................................................. 9 Readout ............................................................................................................................................... 10 Sensor Life .......................................................................................................................................... 10 Active and Passive Sampling .............................................................................................................. 10 Response ............................................................................................................................................. 10 Cost ..................................................................................................................................................... 11 WET METHODS ....................................................................................................................................... 11 Standardized Wet Methods.................................................................................................................. 11 pH Test Paper Methods ...................................................................................................................... 12 GAS DETECTION TUBES .......................................................................................................................... 13 Active sampling tubes ......................................................................................................................... 13 Passive Sampling Tubes...................................................................................................................... 14 FOURIER TRANSFORM INFRARED SPECTROSCOPY ................................................................................. 14 NON-DISPERSIVE INFRARED GAS ANALYZER ........................................................................................ 15 Photo Acoustic Multi-gas Monitor...................................................................................................... 15 Rosemount and Beckman Industrial Models....................................................................................... 16 ULTRAVIOLET DIFFERENTIAL OPTICAL ABSORPTION SPECTROSCOPY .................................................. 17 Opsis AR-500 UV Open-path Monitor................................................................................................ 17 WSU System ........................................................................................................................................ 17 CHEMILUMINESCENCE ANALYZER ......................................................................................................... 18 Matthus-IMAG Converter and Monitor-Labs Analyzer .................................................................... 19 Matthus-IMAG Converter and THIS NOx Analyzer .......................................................................... 19 TEI Converter and Analyzer ............................................................................................................... 19 MT Converter and API Analyzer ........................................................................................................ 20 CHEMCASSETTE DETECTION SYSTEM ..................................................................................................... 20 ELECTROCHEMICAL SENSOR................................................................................................................... 21 Quadscan Gas Monitoring System...................................................................................................... 21 Drger Sensor..................................................................................................................................... 21 Twistik Transmitter ............................................................................................................................. 21

SOLID STATE SENSOR ............................................................................................................................. 21 COMPARISON STUDIES OF MEASURING DEVICES ................................................................................... 22 CALIBRATION OF MEASURING DEVICES ................................................................................................. 23 DISCUSSION ............................................................................................................................................ 25 DEVELOPMENT AND APPLICATION OF MEASUREMENT TECHNIQUES .................................................... 25 DATA QUALITY OF AGRICULTURAL AMMONIA MEASUREMENT ........................................................... 25 SOURCES OF MEASUREMENT ERRORS .................................................................................................... 25 Calibration Gas .................................................................................................................................. 26 High Frequency Variation of Ammonia Concentrations .................................................................... 26 Interference of Water Vapor, Other Gases and Particular Matter..................................................... 28 SELECTION OF SAMPLING TECHNIQUES .................................................................................................. 28 SELECTION OF MEASUREMENT TECHNIQUES ......................................................................................... 29 STANDARDS FOR AGRICULTURAL AMMONIA MEASUREMENT ............................................................... 30 CONCLUSIONS ....................................................................................................................................... 30 ACKNOWLEDGEMENTS ..................................................................................................................... 31 REFERENCES.......................................................................................................................................... 31

Abbreviations and Acronyms


AQT CL DQIs EC IR NDIR OP-FTIR PAS QAQC RIPH TWA USEPA UV UV-DOAS Ammonia Quick Test Chemiluminescence Data quality indicators Electrochemical Infrared Non-dispersive infrared open-path Fourier Transform Infrared Photo acoustic spectroscopy Quality assurance and quality control Research Institute of Pig Husbandry, Rosmalen, The Netherlands Time weighted average United States Environmental Protection Agency Ultraviolet Ultraviolet Differential Optical Absorption Spectrometer

ii

Introduction
Ammonia (NH3) is a common substance playing an important role in the nitrogen cycle. Since the 1980s, agricultural NH3 emission has become one of the major worldwide air pollution problems and has attracted more and more attention from the public and government regulators. It is believed that excessive emissions of NH3 from agriculture to the atmosphere have caused direct and indirect damage to the ecosystem in some regions with intensive animal production (van Breemen et al., 1982; Roelofs and Houdijk, 1991; Slanina, 1994). Van der Hoek (1998) estimated that 8095% of the total NH3 emission in Europe originates from agricultural practices. Their estimates also indicated that ammonia emission from animal excreta contributed over 80% and emissions from use of fertilizers contributed less than 20% of the total NH3 emission. Kurvits and Marta (1998) reported that most NH3 emissions in Canada are from farm animals. The authors also estimated a 21% increase of NH3 emission from animal husbandry in Canada from 1990 to 1995. High concentrations of NH3 inside the animal houses also represent potential health hazards to humans and animals (Reece et al., 1980; Carr et al., 1990; Crook et al., 1991; Wheeler et al., 2000a). Chronic respiratory diseases of swine production facility workers have been attributed to dust and NH3 (Donham et al., 1995). Animal respiratory diseases, such as sneezing, coughing, or pneumonia, increased when NH3 concentrations were 2040 ppm as compared with 515 ppm (Busse, 1993). Although studies of agricultural NH3 have increased in recent years, reliable field measurements of NH3 at animal facilities (animal houses, and manure storage and treatment) are a major need. This is especially true for North America. Kurvits and Marta (1998) explained that the NH3 emission factors used for the Canadian Inventory were taken from the United States Environmental Protection Agency (USEPA), which in turn were based largely on a 1992 Dutch report of emission in the Netherlands. Understanding and control of NH3 at animal facilities depend on sampling/measurement techniques, including devices, instruments, and procedures. Accurate and reliable techniques provide high quality data that are essential to research as well as abatement of NH3 emissions. For many years until now, there have been only a few publications focusing on the measurement of NH3 at animal facilities. Some of the publications reported specific measurement setups. Van 't Klooster and Heitlager (1992) presented a special report describing in detail a system at the Research Institute for Pig Husbandry (RIPH) in Rosmalen, the Netherlands. Berckmans and Ni (1993) described a field installation on a commercial pig farm in Belgium. Heber et al. (2001) provided a detailed description of a field test system including NH3 measurements in eight commercial swine buildings. Ammonia measurement techniques have been reviewed in other publications. Kamin et al. (1979) included, in their comprehensive study of NH3, one chapter on NH3 measurement and monitoring techniques. However, there have been new techniques developed and introduced in agriculture since their book was published. Van Ouwerkerk (1993) reviewed various techniques of NH3 emission measurement at animal houses. McGinn and Janzen (1998) presented measurement techniques to determine the loss of NH3 from manure-amended soils, including micrometeorological techniques and a chamber method, but no measurement of NH3 at animal facilities was discussed. Monteny and Erisman (1998) discussed NH3 measurement in dairy cow buildings including the measurement of ventilation rate in naturally ventilated barns. Ni (1998) reviewed NH3 measurement techniques especially related to animal houses. Phillips et al.

(2001) recently published a comprehensive review of techniques for NH3 emission measurement. However, most of the useful information is still scattered in the literature. Furthermore, technical difficulties in field sampling and measurement need more careful study. A thorough and updated evaluation of NH3 measurement techniques related to animal facilities will benefit users and researchers in selecting, employing, and developing such methods. The objective of this paper was to draw an accurate picture of the state-of-the-art of NH3 sampling/measurement methods/devices based on reported applications at animal facilities, to address technical and practical difficulties in their implementation, and to delineate the future research needs in the area of measurement methodology.

A General View of Ammonia Measurement


To obtain accurate information about NH3 at animal facilities, suitable measurement techniques have to be adopted and one or more measurement variables have to be chosen depending on measurement objectives. These variables include NH3 concentration, air exchange rate (or ventilation rate), air temperature, and air pressure. An example of measurement at animal house is illustrated in fig. 1.

Measurement variables Outdoor concentration Indoor concentration Air exchange rate Air temperature Air pressure Indispensable

Measurement objectives Outdoor concentration

Indoor concentration

Emission from house

Optional

Figure 1: Measurement variables for ammonia concentration and emission at animal building.

To obtain atmospheric NH3 concentrations inside and outside animal buildings, measurement of concentrations at required locations is indispensable while all the other variables are optional, because they are relatively less important. To obtain NH3 emission from animal building or manure storage, the measurement of NH3 concentration difference between the outgoing and incoming air is essential along with the air exchange rate (or ventilation rate). Phillips et al. (2001) provided a comprehensive review of available techniques for measuring building air exchange rates. Most NH3 concentration measuring devices provide direct reading in volumetric concentrations. However, mass concentrations are required to calculate NH3 emissions. The volume of gas depends on temperature and pressure and is therefore not constant. When converting from volumetric concentration to mass concentration, the volumetric concentration is multiplied by the

molecular weight and the pressure, and divided by the gas constant and the temperature. Temperature and pressure therefore need to be known. However, although measurement of air temperature was often included in published works, air pressure measurements are seldom found. Atmosphere pressure varies between about 980 and 1040 mbar, a 6% variation, or a 3% from standard atmosphere, which is often assumed. The measurements of temperature and air pressure are relatively easy with few technical challenges. Reported sampling and measurement methodologies applied at animal facilities are summarized in fig. 2. Details of these techniques are discussed in the following sections.

A M M O N I A

Sampling device / Converter / Collection medium / Measuring device Wet chemistry VECHTA Aqueous Titrimetry acid trap Photometry & colorimetry Ferm tube Conductimetry pH paper Distilled pHydrion test strips water Ammonia Quick Test Gas detection tubes Drger tube Dynamic chamber Kitagawa tube Gastec tube Sensidyne tube MSA tube FTIR spectroscopy K300 M 2401 Open-path Bomen-100 Non-dispersive IR analyzer Stream PAS Type 1302 controller Rosemount; Beckman Dynamic chamber UV-DOAS system Opsis monitor Open-path WSU system Lindvall box CL NOx analyzer NH3 - NO Stream Monitor-Labs analyzer converter controller THIS analyzer Stream TEI analyzer NH3 - NO controller converter API analyzer Electrochemical sensor Flux chamber Drger sensor Quadscan gas monitor Twistik Transmitter Chemcassette Monitor Single Point Monitor Solid-state sensor Solidox sensor IMEC sensor

Figure 2: An overview of techniques applied to measure ammonia concentrations at animal facilities.

Ammonia Sampling
Location and Time of Sampling
Because of the large volume of air flowing through animal houses or flowing over manure storage and treatment facilities, it is impossible to capture all the air for gas concentration determination. Therefore, air must be sampled in order to perform gas concentration measurement. The technical term sampling has different definitions in different research fields. In the case of gas measurement at animal facilities, sampling is defined in this work as: The technique and procedure that specifies the location where air sample is taken, controls the time of measurement (interval, frequency, and duration), and regulates the mass (usually volume) of sample air to be measured. The necessity of selecting location and time of sampling is obvious, because there are temporal and spatial variations of NH3 concentration in animal buildings and open-air manure storage and treatment. An animal building is a ventilated and imperfectly mixed air space, where temperature and concentration gradients exist. Changes in room temperature and building ventilation usually follow diurnal and seasonal patterns. Although ventilation in the building creates air mixing, it can also increase the spatial concentration variations in situations when it dilutes NH3 at some locations more than other locations. Field studies have confirmed nonhomogeneity of NH3 concentrations in livestock houses (Krause and Janssen, 1990; 1991; Berckmans et al., 1994; Ni et al., 2000b). An example of NH3 concentration variations at three different locations during a 24-h period, and the effect of ventilation airflow on the concentrations is provided in fig. 3, which also shows the temporal variation of the NH3 concentrations.

16 Concentration, mg/m
3

250 200 150 Ventilation, 1000 m /h


3

12 8

100 4 WF 0 0 3 6 9 12 15 18 21 24 Time, hour of day PF PH V 0 50

Figure 3. A 24-h record of temporal and spatial variations of ammonia concentrations at wall fans (WF), pit fans (PF) and pit headspace (PH), and ventilation rate (V) in a 66m (L) x 13m (W) mechanically ventilated swine finish building. Source: Ni et al. (2000b). Accurate control of air sample volume is necessary for some analytical techniques (e.g. wet chemistry using acid traps) that only determine total mass of NH3. In order to get volume specific NH3 data (concentration) of the air, the volume of the air sample must be known. Sometime a sampling/control device is needed to perform sampling. However, the

manipulations of space, time, and mass are not always necessary depending on different measurement objectives.

Methods and Devices of Sampling


Closed Sampling: Sampling Chambers Closed sampling methods involve a physical enclosure or a sampling chamber to create a limited headspace over the NH3 release surface that is physically separated from surroundings. A sampling chamber has an open-bottom face and is usually equipped with air inlet(s) and outlet(s). The chamber is placed on the floors of animal buildings or on the surfaces of liquid or solid manure storages that release NH3. Sample air is drawn from the outlet and/or inlet of the chamber. Closed sampling methods provide a small-scale cut-off space under field conditions and therefore enable easily controllable studies. For example, airflow through the chamber can be easily measured and adjusted to investigate its effect on gas release (fig. 4). However, the enclosure may also alter airflow patterns and gas concentration gradients of the sampling headspace that would naturally occur. Thus, there is a significant difference between artificial and natural airflow patterns and gas concentration gradients above the NH3 release surface.

Air stirring fan Flow controller Zero-air Inlet Lagoon

In-situ analysis Outlet Floating platform

Figure 4. Schematic of a closed method used for lagoon ammonia sampling. Source: Aneja et al (2000). Sampling chambers have been given various names, e.g. Lindvall box, dynamic chamber, wind tunnel, hood, convective flux chamber, etc. Their volumes were between 0.02 to 108 m3 and most of them had box-like shapes. Airflow through the chamber was supplied by a fan (installed in the inlet or outlet) or a compressed air cylinder. Most tested involved blowing ambient air, filtered or not, through the chamber, while others blew zero-air (fig. 4). Some chambers utilized a stirring fan inside the chamber to mix air. A summary of the sampling chambers is given in table 1. The earliest sampling chamber used in animal building was a Lindvall Box (1.5 1.0 0.4 m3) developed in Sweden for odor analysis, during which cleaned air was blown through the box and brought back to a mobile laboratory for measurement (Lindvall et al., 1974). The Lindvall Box was later used for NH3 sampling at animal facilities by Scholtens (1990), Kant et al. (1992), Elzing and Swierctra (1993), and Kroodsma et al. (1993). Elwinger and Svensson (1996) described a battery-powered dynamic chamber, developed by Svensson and Ferm (1993), to sample and measure NH3 emissions in an experimental broiler house. Andersson (1995) converted manure culverts in a 160-m2 pig house into a measuring system by covering the slatted floor with boards and placing inlets and outlets of the system. Ferguson et al. (1997) developed an isolated 19-L container laid over a portion of the litter in a

broiler room to sample equilibrium NH3 concentrations every 20 min. A small fan in the container mixed the sample air at a constant velocity (1.0 m/s). Table 1. Summary of the reviewed sampling chambers. Chamber name Lindvall box Dynamic chamber Chamber Container Chamber Dynamic chamber Open dynamic chamber Wind tunnel Hood Shape Box Box Box --Box Box Box --Box Dimension 1.51.00.4 m3 0.400.300.18 m3 L: 5.4m Vol.: 19 L 1.220.760.41 m3 Bottom: 0.25 m2 632 m3 and 962 m3 20.5 m2 Open face: 0.9 m2 1.220.610.25 m3 Incoming air Cleaned Ambient Ambient None Ambient (0.2m/s) Ambient Ambient (111 m3/h) Ambient (1.0m/s) Filtered (1.0 m/s) Filtered (1.1 m/s) Emission Stir fan Ref source Animal houses Broiler house Pig house Broiler litter Broiler litter Pig house Treated manure Concrete floor Dairy farm Lagoon Lagoon No No No Yes No Yes No No No No Yes 1 2 3 4 5 6 7 8 9 10 11

Buoyant Convective Box Flux Chamber Dynamic chamber

Cylinder 0.27m()0.42m(h) Zero-air (2.44.7 L/min)

Notes: 1. Scholtens (1990), Kant et al. (1992), Elzing and Swierctra (1993), Kroodsma et al. (1993); 2. Elwinger and Svensson (1996); 3. Andersson (1995); 4. Ferguson et al. (1997). Shape of the chamber was not described; 5. Brewer and Thomas (1997); 6. Svensson et al. (1997); 7. Amon et al. (1997); 8. Misselbrook et al. (1998). Shape and height of the wind tunnel were not given by the authors; 9. Misselbrook et al. (1998); 10. Heber et al. (2000a); 11. Aneja et al (2000).

Brewer and Thomas (1997) used a Plexiglas chamber to cover the litter in a commercial broiler house. Svensson et al. (1997) employed a dynamic chamber combined with diffusion samplers and gas detection tubes in a naturally ventilated pig house. Air from 2.0 m above the bedding surface was sucked into the chamber by a fan. An air-mixing fan was installed inside the chamber. Amon et al. (1997) described two large-scale open dynamic chambers developed in Austria to collect emission data. The incoming air to the chamber was fresh ambient air. The chamber was used to study gas emissions from aerobically composted and anaerobically stored solid manure. Misselbrook et al. (1998) conducted sampling on concrete floors and a commercial dairy yard using a wind tunnel and a hood, respectively.

Heber et al. (2000a) developed a buoyant convective flux chamber for odor and gas sampling using 1.0m/s air speed in the chamber. It floated about 0.17 m above the water and covered 0.74 m2 of swine lagoon surface. The air to the chamber was provided by an air supply unit that contained activated charcoal, zeolite and permanganate. Aneja et al. (2000) used a 27 cm diameter, 42 cm height open bottom cylinder with a Teflon-coated wall and a stirring fan. The cylinder was inserted into a buoyant platform and was ventilated with compressed zero-grade air. It was used to sample NH3 from lagoons (fig. 4). Point Sampling and Time Control: Air Stream Controller Single- or multi- point sampling selects one or more point locations where sample air is diffused to NH3 samplers or sensors, or is pumped to NH3 measuring devices. This sampling method does not disturb the NH3 source and its surroundings. It is the most widely used sampling method, especially when using measuring devices such as detection tubes, because of its simplicity. When multi-point sampling is needed with a single measuring device, an air stream controller can be used to regulate the time and location of sampling by controlling selected air stream(s) from multiple locations, usually transported in tubing, to the measuring device. Computer aided stream controllers enable automation and efficient use of limited numbers of expensive analytical instruments. An air stream controller placed between several NH3 NO converters and a NOx analyzer was described by van Ouwerkerk (1993). Berckmans and Ni (1993) employed a similar device that regulated air sample streams two minutes for each of six converters. Neser et al. (1997) used an air stream controller to scan through four sampling points including three laying hen compartments and one incoming air location, 15 min at each sampling point. Heber et al. (2000b; 2001) and Ni et al. (2000a; 2000b) described an air stream controller to regulate air streams from up to six sampling locations, each lasting 10 to 15 min, to a single NH3 NO converter that was followed by an NH3 analyzer (fig. 2). Open-path Sampling: Optical Detection Device Open-path sampling uses optical detection device, which consists of an emitter telescope and a receiver/detector. The source light from the emitter, ultra violet (UV) or infrared (IR), is beamed in one direction through a space (hence an open path), which contains gaseous NH3, to the receiver/detector (fig. 5). Depending on the technology, the length of the open-path between the light source and receiver can be several meters, 100150 m (Secrest, 2000), up to 500 m (Amon et al., 1997) or 750 m (Mount et al., 2001). The source and the receiver can be at the two ends of the open-path. The detection path is between the source and the receiver. They can also be at the same end with a reflection mirror at the other end of the open-path. In this case, the light source beamed to the distant mirror reflects back to the receiver. The path is between the mirror and the source/receiver. The open-path technique provides path-averaged gas concentrations. Open-path sampling has a shorter history than point sampling for agricultural NH3. One advantage of the open-path technique is that it does not disturb the system being measured. Also no adsorption of NH3 or other gases on parts of the measuring device is possible. Large areas can be investigated and the detection limit is very low. Disadvantages of the method lie in the determination of emission rates by the inverted dispersion models. Climatic conditions must be known during the measurement period. Different emission sources lying close to each other cannot be distinguished from each other (Amon et al., 1997).

Light source Open-path Detector

Ammonia source

Figure 5. Schematic of open-path sampling. According to Amon et al. (1997). Open-path sampling for NH3 concentration was conducted outside an animal building for emission rate assessment. Secrest (2000) oriented the monitoring paths east-west when the prevailing winds were from the south at 0.8 km north of a cluster of eight swine barns in Missouri using an Ultra Violet Differential Optical Absorption Spectrometer (UV DOAS) and a Fourier Transform Interferometer. Harris et al. (2001) employed a single-beam open-path Fourier Transform Infrared (OP-FTIR) system along several downwind paths to measure all the exhaust plumes from nine finishing swine barns in North Carolina. The OP-FTIR beam passed through the fan plume one meter from the fan outlet. Mount et al. (2001) measured NH3 emission on a dairy farm with a path length up to 750 m. Sampling Air Volume Control Pfeiffer et al. (1993) and Krieger et al. (1993) adopted an air collection system called VECHTA to control sampling volume. The system continuously provided 1.0 L/min of sample air to an acid collection medium, but the authors did not give detailed information about this device. Phillips et al. (2000) utilized a revised Ferm tube, a passive sampling device, in a large dairy cow house. The essential feature of this sampler was that a precision orifice in a disc of very thin material was installed in the mid-section of the tube so that the air flowing through the tube was proportional to the wind speed. Any NH3 passing through the sampler was captured by a coating of oxalic acid deposited on the inside wall of the tube by the evaporation of a solution in acetone. Like other acid trap samplers, captured NH3 after exposure of the sampler was tested using colorimetry or other analytical methods. Jacobson et al. (1998) collected air samples in 10-L Tedlar bags each week from the ventilation exhaust air and inside the barns for measurement of NH3 concentration using colorimetric tubes. Heber et al. (2000a) used 50-L Tedlar bags for lagoon air sampling. However, while these bags defined the sample air volume, the main purpose of their use was for sample storage and/or transportation.

Ammonia Concentration Measuring Devices


Classification
Some characteristics of the techniques summarized in fig. 2 can be used to classify features to be considered during the selection process. However, the characteristics of different techniques are all relative to each other. Classifications are made in this paper according to analytical methods (wet or dry), sensitivity, delivery of measurement results (direct or indirect readout),

type of sensor use (single or multiple use), method of sampling air delivery (active or passive sampling), response time, and cost of the device (table 2). Table 2. Characteristics of some reviewed ammonia measuring devices. Wet or dry Standardized wet chemistry pH paper/test strip Gas tubes Passive gas tubes Passive sampler EC sensors NOx analyzers Wet Wet Dry Dry Dry Dry Dry Sens. 1 0.011 mg/L ppm ppm ppm ppb ppm ppb ppb ppm 0.011 ppm ppb ppm ppm Point or path Point Point Point Point Point Point Point Path Point Point Path Point Point Readout Indirect Direct Direct Direct Indirect Direct Direct Direct Direct Direct Direct Direct Direct Sensor Active or Resp. 4 Cost 5 use 2 passive 3 S S S S S M M M M M M S M A P A P P A; P A P A A P A P s s s 35 s
9

h s min h h 2 min

L6 VL L L L H VH VH H-VH VH H H M

FTIR spectroscopy Dry Rosemount NDIR 7 Dry PAS 1302 UV-ODAS Chemcassette Solid state sensor
8

Dry Dry Dry Dry

Notes: 1 Sensitivity, expressed in concentration level; 2 S, single use. M, multiple use; 3 A, active. P, passive; 4 Response time: h, hours. min, minutes. s, seconds; 5 Costs are approximate and may vary depending on time and location of purchase, providers and accessories. Cost is investment only; maintenance and operation are not included. VL, very low (<$1/sample). L, low ($1 10/sample). M, medium (<$5000/device). H, high (>$5000 10000/device). VH, very high (>$10000/device); 6 Analytical instrument used to analyze the collection medium is not counted; 7 Rosemount Non-dispersive Infrared Analyzer; 8 Photoacoustic Spectroscopy Type 1302; 9 For one gas or vapor if the tube is shorter than one meter.

Wet and Dry According to Kamin et al. (1979), analytical methods of NH3 can be classified as wet method, which uses an aqueous medium, and dry method, which is the method of direct analysis of NH3 in the gas phase. Sensitivity Sensitivity is the capability of a measuring device to discriminate between measurement responses representing different levels of a variable of interest. Sensitivity is determined from the value of the standard deviation at the concentration level of interest. It represents the

minimum difference in concentration that can be distinguished between two samples with a high degree of confidence (USEPA, 1998). Some of the techniques were reported to be highly sensitive, for example, the method of converting NH3 to NO followed by NOx analysis is sensitive at the ppb level. Other techniques provide sensitivity at ppm level, e.g. NH3 detection tubes. Sometimes high sensitivity techniques are called analytical techniques, which provide quantitative data, and low sensitivity ones are called detection techniques, which provide qualitative or semi-quantitative data. Readout Direct readout is an important feature, especially for field measurements. Techniques with direct readouts provide an immediate visual display of NH3 concentration right after the measurement is completed. Most of the reported techniques provide direct readout. Some of them are followed by automatic data retrieving and processing. Techniques with indirect readout require a chain of procedures and devices; for example, trapping NH3 in acid collection medium followed by laboratory analysis of the medium with wet chemistry methods. Compared with the direct readout, the indirect readout method takes more time to obtain results and is not suitable for large numbers of measurements. Sensor Life The sensor in an NH3 measuring device is the material or part that undergoes a physical or chemical change when exposed to sample air. Single measurement techniques adopt disposable sensor materials that cannot be re-used. The gas detection tube is a typical single use sensor. In wet chemistry methods, the NH3 collection medium is used only once. Continuous measurement techniques can provide many concentration readings over time, usually in the form of electronic signals that can be easily recorded and processed. Sensor materials for these devices usually do not need to be replaced, with the exception of EC sensors (e.g. Drger unit that has a sensor life of 18 months). The cassette needed with the Chemcassette Gas Monitor is a single use sensor, although one cassette can provide multiple readings. Active and Passive Sampling An active sampling technique needs a pump, whether hand or electric powered, to force the sample air flowing to the sensor. Techniques with active sampling also enable transportation of sample air through sampling tubing to realize multi-point measurements with a single set of measuring device. A passive sampling technique lets air diffuse into the sensor, or lets air stay as is in the open measurement point, thus a pump is not required. Sensors for passive sampling need to be placed right at the sampling location during measurement. Passive sampling depending on diffusion takes longer to finish a measurement. Because of this, it can provide time-weightedaverage (TWA) concentration in a single point measurement. Some active methods (e.g. wet chemistry using acid traps) also provide TWA concentrations. Response Response is a measure to evaluate how quickly or slowly a measuring system can react to NH3 and present correct concentration readings. Manufacturers use response time (TEI, 1995), time needed for an instrument to reach from 0 to 90% at zero to span difference in gas

10

concentrations, to describe specifications of the instrument. The response of passive diffusion sampling can be as long as 8 or 24 h. Cost Costs of the reported techniques vary widely. The lowest of single-measurement is < $0.10 (e.g. pH paper test strips) if only sensor material is counted. A gas detection tube may cost about $5.00. The most expensive measuring device with a multi-use sensor can be >$20,000 (e.g. the PhotoAcoustic Spectroscopy (PAS)). Cost may also be much different if accessories (e.g., calibration kit, stream controller, etc.) are added to the instrument.

Wet Methods
Standardized Wet Methods Most wet methods are standardized methods (table 3). With wet methods, once valid samples of ammonium ions are obtained in solution, it is relatively simple to arrive at final analytical results in the laboratory. Table 3. Standardized wet methods used for measuring ammonia in acid traps. Methods ColorimetryNessler Sensitivity 0.02 mg/L Comments Traditional method, widely used in past; Numerous interferences by other compounds, including aldehydes, sulfur dioxide, amines, and metals; Pre-purification by distillation often recommended. Widely used; Adapted for automated analysis; Less sensitive to interference than Nessler method; pHdependent. Relatively simple but cannot handle large amount of samples. NA Potential interference from other redox species. All acids and bases interfere. Ref. 1

Colorimetryindophenol Photometry PhotometryNessler Conductimetry Titrimetry

0.01 mg/L

NA NA 0.1 mg/L 1 mg/L

3 4 5 6

Notes: Some references cited in this table provided technical information of the methods but did not report NH3 measurement in animal houses. NA: not available from the cited references. References: 1. Valentine (1964), Hashimoto (1972), Kamin et al. (1979), Valli et al. (1991), Kay et al. (1992); 2. Kamin et al. (1979), Hesse (1994); Hoy et al. (1994); 3. Leithe (1971), Schmidt-Van Riel (1991), Wang et al. (1991); 4. Jiang and Sands (2000); 5. Kamin et al. (1979), Mannebeck and Oldenburg (1991); 6. Curtis et al. (1975), Verstegen et al. (1976), Kamin et al. (1979).

These methods rely on collecting gaseous NH3 into a suitable acid solution (acid trap, or scrubber) and then performing concentration determination (fig. 6). The volume of air passed through the scrubber is recorded and the NH3 concentration in the air is calculated (Hashimoto, 1972).

11

The most commonly used acid traps for measuring NH3 at animal facilities include boric acid (H3BO3) (Curtis et al., 1975), sulfuric acid (H2SO4) (Valli et al., 1991; Krieger et al., 1993, Guingand, 1997 #1681; Pfeiffer et al., 1993; Jiang and Sands, 2000), and orthophosphoric acid (Asteraki et al., 1997; Kay and Lee, 1997; Misselbrook et al., 1998). Three general methods of colorimetric techniques have accounted for the majority of practical use to determine the NH3 concentration in the acid traps. They are Nessler, Indophenol, and pyridine-pyrazolone techniques; each of them has modifications and adaptations. However, only the first two, the Nessler and Indophenol techniques, have been applied to the NH3 measurement at animal facilities. Photometry constitutes one of the most important methods in air analysis; the air samples must be converted into colored compounds that are then determined. There are specific and very sensitive color reactions available and the time required to perform a comparatively accurate measurement is short. Jiang and Sands (2000) analyzed NH3 concentration in broiler buildings using Nessler method and a spectrophotometer. Non-colorimetric wet-chemical techniques include acid conductimetry and titrimetry. Conductimetry was used for measuring NH3 emission from an animal house (Mannebeck and Oldenburg, 1991). Generally these techniques tend to be less sensitive than the colorimetric method and are subject to a host of interferences (Kamin et al., 1979). Titrimetry was also used for measuring NH3 in animal houses (Curtis et al., 1975; Verstegen et al., 1976). It is relatively simple and inexpensive, but is less sensitive and is subject to interference.

Sample air

Flow meter

Pump

Exhaust

Volume control

Acid solution

Concentration determination

Figure 6. Schematic of sampling and analysis with wet chemistry. pH Test Paper Methods Moum et al. (1969) developed a very simple method by employing pH test paper and neutral distilled water as an NH3 trap. The measuring range was 0100 ppm and the accuracy was 5 ppm. One follow-up use was tested by Seltzer et al. (1969). The method is inexpensive and provides direct in-situ readout. However, it has low sensitivity and precision. Using a commercialized version of this method, Dewey et al. (2000) tested pHydrion NH3 test strips (Micro Essential Laboratory, Brooklyn, NY) in a swine unit in 1994. The pHydrion cost only $0.06 per test. It involved placing a drop of distilled water on a paper test strip, waving the strip in the air for one minute, and estimating NH3 concentrations by matching the color change with a calibrated color chart. A similar product, Ammonia Quick Test (AQT) , distributed by Vineland Laboratories (Vineland, NJ), has a measurement range of 0100 ppm NH3. Skewes and Harmon (1995) tested the AQT

12

in eight commercial broiler houses and concluded that it estimated NH3 levels accurately at 20 25 ppm levels.

Gas Detection Tubes


Gas detection tubes are based on adsorption of tested air pollutant on solid surfaces accompanied by a color reaction. There are two types of disposable tubes: active sampling tube and passive sampling tubes. Tubes with different measurement ranges are available. Some have suitable measurement ranges for NH3 in animal buildings. Usually, the sensitivities of the tubes are too low for measuring outdoor NH3 concentrations. The most obvious advantage of the gas detection tube is its operational and functional simplicity. Therefore, it was widely used in agricultural NH3 measurement. Meyer and Bundy (1991) measured NH3 concentrations in 200 pig farrowing houses with gas detection tubes. Gas tubes are relatively low cost, usually $5 to 8 per tube and $300 to 500 for a hand pump. Dewey et al. (2000) reported $306.00 for the hand pump (Drger Accuro Pump, Drger Safety Inc., Pittsburgh, Pennsylvania) and $4.12 for each tube. Leithe (1971) and Leichnitz (1985) stated that the standard deviation of the results of Drger gas detection tubes is about 68% in favorable and 1020% in less favorable cases. Skewes and Harmon (1995) found that the passive tubes (Gastec Passive Dosimeter Tube No. 3D) estimated average NH3 levels accurately at low levels of NH3 as compared with the Gastec Ammonia Low Range Detector Tube No. 3La. Liu et al. (1993) concluded that the accuracy of NH3 detector tubes from Mine Safety Appliance Company in Pennsylvania was 1 ppm. According to Scholtens (1993), Drger tubes Type 2A and Kitagawa tubes Type 105SD for NH3 measurement both had precision (coefficient of variation) less than 5%, and an inaccuracy (systematic error) less than 10%. The precision of Drger gas detector tubes (Type 5A) was approximately 10% and the inaccuracy was about 15%. Both precision and inaccuracy of Kitagawa tube (Type 105SC) were approximately 2%. The precision and inaccuracy of the gas tubes became worse at lower NH3 concentration levels. Active sampling tubes Active sampling involves a hand-pump that sucks a pre-defined volume of air per stroke (fig. 7). Both ends of the test tube are sealed when manufactured and are cut open just before measurement. The open-end tube is inserted tightly into the pump connector. By pumping the hand-pump, the air sample flows through the tube. The color that arises is evaluated to assess the NH3 concentration. Active gas tubes from five different manufacturers have been used for NH3 measurement at animal facilities. The Drger Tube is probably the most widely used gas tube product. Patni and Clarke (1991) used Drger tubes to measure short-term NH3 concentrations in a dairy cattle barn, a heifer barn, swine barns and caged-layer barns. Scholtens (1993) tested and compared the Drger tubes with the Kitagawa tubes. Stowell and Foster (2000) and Stowell et al. (2000) reported using Drger tubes to sample NH3 in exhaust air from three selected fans of a 960-head HighRise swine building. The use of Kitagawa gas detector was reported by Reece et al. (1979), Johnston et al. (1981), Scholtens (1993) and Svensson et al. (1997). De Praetere and van Der Biest (1990), and Jacobson et al. (1992) reported the measurement of NH3 using Gastec Tubes. Xin et al. (1996) used Sensidyne detector tubes (Sensidyne, Clearwater, FL) to check NH3 concentration at bird level in poultry buildings.

13

Wheeler et al. (1999) tested the Kwik-Draw Pump (Mine Safety Appliances Company, Part number#487500) with MSA detector tubes (MSA #804405), which was designed for one-hand operation and consistent delivery of a 100-mL sample volume. An integral counter helps track number of sample pump strokes.
Gas tube Concentration scale Tube connector Hand-pump

Direction of sample airflow and tube color change Chain

Figure 7: Drger gas test tube and hand pump Passive Sampling Tubes Like the active tubes, the passive sampling tubes are also sealed before using. However, only one sealed end of passive tube is broken open to commence measurement. The opened tube is exposed at the selected sampling location for a specific time, usually several hours. The gas concentration indicated in the tube should be interpreted with the exposure time. Passive NH3 sampling tubes from different manufacturers have also been used at animal facilities. Patni and Clarke (1991) used Drger tubes for TWA NH3 concentrations in dairy, swine, and poultry barns. Busse (1993) used a passive Drger system with a diffusion tube and a holder, in which the tube was inserted for 24-h measurements. Nicks et al. (1993) reported using 8-hour diffusion tubes. Nicks et al. (1997), Choinire et al. (1997), and Wheeler et al. (1999) employed Gastec diffusion tubes. Pratt et al. (2000) used diffusion tubes in horse stalls in Kentucky, but did not indicate the tube manufacturer.

Fourier Transform Infrared Spectroscopy


Fourier Transform Infrared (FTIR) spectroscopy is a technique involving the interaction of IR electromagnetic radiation with the test sample. The technology has been called interferential spectroscopy; multiplex; Fourier spectroscopy; interferometric spectrophotometry, or Fourier transform spectroscopy through its development by physicists and manufacturers over the years. The acronym FTIR is almost universally used by chemists to refer to the technique (Johnston, 1991). The Fourier transformation is a mathematical manipulation that relates a signal, curve, or algebraic function to its frequency content. In Fourier spectroscopy, the output signal is known as an interferogram, and is produced by an interferometer (fig. 8). Interferometers used in FTIR instruments manufactured in recent years are similar in design to the one built at the end of 19th century (White, 1990). As the movable mirror is gradually displaced, a cycle of maximum and minimum intensity recurs. It yields specific information about the chemical structure of organic and inorganic compounds based on the unique vibrational modes of different chemical bonds. The FTIR spectrum is rich with information because each vibrational mode absorbs a specific wavelength of IR radiation. Each bond within a molecule may have several vibrational modes. The FTIR absorption spectrum is a "fingerprint" for a particular molecule that can be compared with reference spectra of known compounds, thereby aiding in the identification of unknowns and providing unambiguous confirmation of the identity of "known" materials.

14

Fixed mirror Light source

Moving mirror Sample Detector

Beam splitter

Figure 8: Schematic diagram of interferometer. Drawn according to White (1990) and Johnston (1991). A few of the many models of FTIR spectroscopy have been used for NH3 measurement at animal facilities. An FTIR Spectroscope K300 with a White-Cell was used in Germany (Neser et al., 1997). Air samples were pumped into or through a special optic cell (White-cell), which used a series of mirrors to create a lengthened light path of 8 m (Amon et al., 1997). A Midac model M2401 FTIR spectroscopy (Harris, personal communications) was used in an outdoor open-path measurement of NH3 emission from swine buildings in North Carolina (Harris et al., 2001). An FTIR ETG w/ Bomen-100 interferometer and sterling cycle cooled detector was used in Missouri (Secrest, 2000; 2001). Other uses of FTIR spectroscopy included an NH3 gas analyzer in the IR/VIS and UV spectral regions (Keck et al., 1994), an IR spectrometer and a data logging system measuring NH3 emission from pig and dairy barns in Germany (Jungbluth and Bscher, 1996; Hartung et al., 1997; Jungbluth et al., 1997), and two other reports by Hauser and Flsch (1993) and Gallmann and Hartung (2000) without technical details.

Non-Dispersive Infrared Gas Analyzer


Non-dispersive infrared (NDIR) analyzers measure the spectral absorption of a gas at one spectral band of the IR spectrum. The spectral dispersion of the absorption spectrum of the gas is not used (Phillips et al., 2001). Photo Acoustic Multi-gas Monitor In Photo Acoustic Spectroscopy (PAS), the gas to be measured is irradiated by intermittent light of pre-selected wavelengths. The gas molecules absorb some of the light energy and convert it into an acoustic signal, which is detected by a microphone. The general principle of the PAS system (Innova AirTech Instruments A/S, Ballerup, Denmark) is illustrated in fig. 9. The PAS monitor can automatically measure multiple gases with a single instrument. When gas samples are drawn from ambient air around the analyzer, the measurement time is approximately 30 s for one gas or water vapor, and about 120 s if five gases plus water vapor are measured. Increasing the length of the sampling tube increases the time required to pump in a new air sample and therefore increases the measurement time. The PAS requires less frequent calibration as compared with NOx analyzers. However, its investment is relatively high and it is subject to interference of water at high relative humilities.

15

Infrared source

Chopper wheel

Optical filters

Analysis cell

Microphone 1 Outlet valve

Air outlet

Flush valve Inlet valve Mirror Optical filter carousel Optical window Microphone 2 Air inlet

Figure 9: Schematic diagram of the general principle of Photo Acoustic Multi-gas Monitor (Type 1302). Courtesy of Innova AirTech Instruments A/S, Denmark A NDIR PAS (Type 1302, Brel & Kjr, Nrum, Denmark) was used intensively in the Netherlands by the RIPH in different studies (van 't Klooster and Heitlager, 1992; den Brok and Verdoes, 1997; Hendriks and Vrielink, 1997; Verdoes and Ogink, 1997; Osada et al., 1998). Use of PAS in Germany included Hoy (1995), who measured five gases quasi-continuously, Snell and Van den Weghe (1999), who used it in a pig building, and Brunsch (1997), who measured gas concentrations in a poultry farm by combining a Type 1302 monitor with a multipoint sampler Type 1309. The PAS has been used in the USA by Ferguson et al. (1997) in a broiler house and by Pratt et al. (2000) in horse stalls in Kentucky. Rosemount and Beckman Industrial Models According to Rosemount Analytical (Orrville, OH), the Model 880A Analyzer produces IR radiation from two separate energy sources. A chopper modulates this radiation into 5 Hz pulses. Depending on the application, the radiation may then pass through optical or gas filters to reduce background interference from other IR-absorbing components. Each IR beam passes through a separate cell. One cell contains a continuous flowing sample while the other cell is either sealed or contains a continuous flowing reference gas. A portion of the IR radiation is absorbed by the component of interest in the sample, with the quantity of IR radiation absorbed being proportional to the component concentration. The detector is a gas microphone based on the Luft principle. It converts the difference in energy between sample and reference cells to a change in capacitance (fig. 10). Brewer et al. (1997) used a Rosemount Analytical Model 880A gas analyzer with a dynamic chamber to measure NH3 concentration in a broiler house in Arkansas. Hartung et al. (2001) described an IR gas analyzer Model BINOS-IR-2 (Rosemount GmbH & Co.), but did not provide details. Maghirang et al. (1991) and Maghirang and Manbeck (1993) used Beckman Industrial Models 770 and 780 (Beckman Industrial Corporation, La Habra, CA) NDIR analyzers in a commercial egg laying house in Pennsylvania. The Beckman models have been discontinued.

16

Infrared sources Chopper

Component of interest Other molecules Air sample in

Reference cell

Sample cell Air sample out

Detector Signal Diaphragm distended

Figure 10: Schematic of operation principle of non-dispersive infrared analyzer Model 880A. Source: Product Date Sheet PDS 103-880A.A01, Rosemount Analytical.

Ultraviolet Differential Optical Absorption Spectroscopy


In the ultraviolet differential optical absorption spectroscopy (UV DOAS) method, an emitterreceiver set creates a light path in a detection zone. Light is generated by a xenon lamp in the emitter and projected to the receiver. Each gas in the detection path absorbs different parts of the light spectrum in a unique way. The absorption is recorded using a spectroscope. Opsis AR-500 UV Open-path Monitor The Opsis AR-500 open-path monitoring system (Opsis Inc., San Marcos, CA) is designated by the USEPA as an Equivalent Method for measuring the criteria pollutants SO2, NO2, and O3 in ambient air. It consists of a light source emitter, a target gas cell, a receiver, a fiber optic cable and an analyzer. The fiber optic cable connects the receiver and the analyzer (fig. 11). According to an environmental technology verification test (Myers et al., 2000), the AR-500 had an NH3 detection limit between 2.8 and 5.8 ppb with good linearity. Its relative accuracy was 3.3 to 11% over the range of 24 to 200 ppb. Secrest (2000; 2001) used a UV DOAS Model AR500 (Opsis AB, Furulund, Sweden) with a Czerny-Turner spectrometer and "ER-110" series (110 mm optics) telescopes to measure ambient NH3 concentration on swine farms in Missouri and Maryland. WSU System An open-path measurement system was developed by Washington State University and has been used for over 10 years in challenging field conditions. Mount et al. (2001) described its use on a dairy farm, where NH3 was measured in the UV bands near a wavelength of 210 nm at an integration time of several seconds to an accuracy of approximately 20%. The system was conceptually simple and consisted of 1) a UV light source, 2) a telescope to beam the UV light into the atmosphere, 3) a mirror system to reflect the light back towards the light source, 4) a receiver telescope to focus the light spectrally absorbed by the atmosphere onto a dispersing spectrograph, 5) a multi-element multiplexing digital detector, and 6) a data analysis system. The path length ranged from a few meters to 750 m and the sensitivity limit was 1 ppbv (parts per billion based on volume).

17

Target Gas Cell Light Source Open Detection Path Fiber Optic Cable Analyzer
Figure 11. Schematic of Opsis monitoring system. Drawn according to: Myers et al. (2000) and http://www.opsis.se/american/am-cem-tech.htm

Receiver

Chemiluminescence Analyzer
Chemiluminescence (CL) NH3 analyzers involve an indirect measurement of NH3 based on converting NH3 to nitric oxide (NO) and then performing NO analysis by CL method. The NH3 content is obtained by either chemical or mathematical subtraction of the background NO signal (Pranitis and Meyerhoff, 1987). This technique requires two instrument modules, an NH3 converter and an NOx analyzer. The converter is kept at a temperature of 795 C and stainless-steel is used as the catalytic active metal. At this temperature, NH3 is converted into NO by the following reaction (Aneja et al., 1978):
4NH3 + 5O2 4NO + 6H2O Fe2O3 CrO3

(1)

The gas phase reaction of NO and ozone (O3) produces NO2 and a characteristic luminescence in the analyzer (fig. 12). When electronically excited NO2 molecules decay to lower energy states, light emission occurs: NO + O3 NO2 + O2 + hv (2)

Pressure in the NO2 detection chamber is kept at least 31 kPa below atmospheric pressure and temperature is kept at 50 C. Under these conditions, the NO concentration is directly proportional to the photon emission intensity. During transport of gas in tubes from the converter to the NOx analyzer, some NO may oxidize to NO2. A molybdenum converter at 325 C converts any NO2 into NO prior to entering the reaction chamber of the NOx analyzer (van 't Klooster and Heitlager, 1992). The advantages of this technique include high sensitivity (1 ppb), high precision (0.5 ppb), linearity (1% full scale), and automation (TEI, 1995). The disadvantages include highly priced parts (e.g., internal gas scrubbers) by some manufacturers, and relatively large initial investment and complicated maintenance. The CL method has been used for NH3 measurement at animal facilities in the Netherlands, UK, Belgium, and the USA.

18

NO detection pathway NO2* detection chamber Sample air NO2 + NO NO2 NO Converter NOx as NO O3 Ozone generator O3 NO2* detection chamber PMT Exhaust [NO2] = [NOx] - [NO] NO signal PMT [NO]

NOx signal

[NO]

NOx detection pathway

Figure 12: Schematic diagram of the general principle of chemiluminescent detector for nitrogen dioxide and nitric oxide. PMT, photo-multiplier tube. Source: Stern et al. (1984).

Matthus-IMAG Converter and Monitor-Labs Analyzer Van 't Klooster et al. (1992) Verdoes and Ogink (1997), and den Brok and Verdoes (1997) used NH3 NO converters type Matthus-IMAG and a dual-channel NOx analyzer (Monitor-Labs Model 8840) in the RIPH at the Netherlands. Scholtens (1990), Aarnink et al. (1993), and Groenestein et al (1997) used an instrument of the same model in the Institute of Agricultural Engineering, the Netherlands. Once a week the monitor was calibrated with a gas of 40 ppm NO in N2 and the flow of the different channels was checked. Dust filters were changed when necessary. A similar system was reported by Demmers et al. (1999) in the U.K. Matthus-IMAG Converter and THIS NOx Analyzer Berckmans and Ni (1993) described another system in Belgium that had five NH3 NO converters (Type Matthus-IMAG) installed at five different sampling locations. Through some heated tubing and a stream selector, the converted air from all the converters was conducted to a single NOx analyzer (Model 42-I, Thermal Instrument System, Franklin, MA (currently Thermo Electron Corporation)). The converters had conversion efficiencies between 95 and 99% at NH3 concentrations below 30 ppm according to the manufacturer. However, one of the five converters was found unstable (Ni, unpublished). TEI Converter and Analyzer Heber et al. (2001) reported four NH3 analyzers (Model 17C, Thermal Environmental Instruments, Inc., Franklin, MA (currently Thermo Electron Corporation)) used in a comprehensive field study of gas emission measurement in Indiana and Illinois. Each NH3 analyzer consisted of two separate modules, a converter module and an analyzer module. The difference between these systems and the systems in Europe is that only one converter was used with each analyzer. An air stream controller was installed before the converter to facilitate multi-point sampling (fig. 2). Several papers on NH3 emissions using these analyzers have been published (Heber et al., 1997a; 1997b; 2000b; 2001; Ni et al., 2000a; 2000b).

19

Harris et al. (2001) and Walker (personal communication) reported using two TEI analyzers (Model 17C) in tunnel-ventilated swine finishing houses in North Carolina. At each house, one analyzer was dedicated to the primary exhaust fan and the other was periodically moved from fan to fan in order to ascertain variability in exhaust NH3 concentrations. McCulloch et al. (unpublished) tested three TEI Model 17C analyzers and found that two of them exhibited excellent system linearity across the 01000 ppb range. In a third analyzer, the converter efficiency varied from 54 to 77% at 12 and 46 ppb, respectively. MT Converter and API Analyzer When measuring lagoon NH3 emission in North Carolina, Aneja et al. (2000) transferred sampling air to a Measurement Technologies 1000N stainless steel NH3 converter. The sample flow from the converter was routed to a CL NOx analyzer (Model 200, Advanced Pollution Instrumentation, San Diego, CA).

Chemcassette Detection System


The colorimetric principle (fig. 13) is employed by the Chemcassette Detection System (Zellweger Analytics Inc., Lincolnshire, IL). A carefully prepared reel of porous paper tape is impregnated with a chemical. The paper acts as both a trapping and analysis medium, detecting and measuring nanogram amounts of target gas. Upon exposure to the target gas, the paper tape changes color in direct proportion to the sample gas concentration. A photo-optical system measures its color intensity change and determines the sampled gas concentration.

Chemcassette Cable

Exhaust air

Photo-optical detector Tape

Sample air
Figure 13. Schematic of ammonia measurement with Chemcassette Monitor.

The system can measure different gases with different Chemcassettes, which are individually formulated for a specific gas or family of gases. The low-level NH3 detection Chemcassette has a range of 0.530.0 ppm. The system measures gas concentrations continuously and the response speed is several seconds. A Single Point Chemcassette system costs $5,200 and an NH3 cassette tape costs $46. According to the manufacturer, this system is the only gas detection method providing physical evidence of gas presence and the technique virtually eliminates in-field calibration. Bicudo et al. (2000) described a Zellweger MDA single Point continuous air monitor at animal facilities in Minnesota. The accuracy of the instrument was 20 % of the actual reading.

20

Electrochemical Sensor
Electrochemical (EC) NH3 sensors consist of two electrodes and detects NH3 with the following EC reactions: on the measuring electrode,
2 NH3 N2 + 6 H + + 6 e and on the counter electrode, 3/2 O 2 + 6 H + + 6 e 3 H2 O . Electrochemical sensors provide direct readout and continuous measurements. Several EC sensors have been tested or used at animal facilities. Quadscan Gas Monitoring System Hoy et al. (1992) and Hoy and Willig (1994) described a Series 6004 Quadscan Gas monitoring system, connected with a printer, for continuous measurement of NH3. This system consists of two main components, the Series 6004 gas receiver and three Series 4485 NH3 gas transmitters. The Series 4485 was a two-wire transmitter designed for monitoring ambient NH3 gas concentration. The transmitter itself consisted of two components, an EC sensor, and an electronic transmitter. The standard 4485 had a measurement range of 0100 ppm. The authors did not provide details of the gas receiver or the EC sensor. Drger Sensor Heinrichs and Oldenburg (1993) used a Drger apparatus for measuring NH3 in a fattening pig house, but did not provide details of the system. Another Drger apparatus, with a measurement range of 1100 ppm of NH3, was tested in a field installation in Belgium (Ni, unpublished). A Drger NH3 sensor (Polytron 2, Drger Safety, Inc. Pittsburgh, PA) connected to a data logger was used in broiler houses (Wheeler et al., 1998; 1999; 2000a). The sensor was battery powered. Its scale was 0300 ppm and its precision was 3% or 9 ppm. The sensor was successful at a reasonable cost for research or demonstration projects. The unit consisted of a multi-gas body (Polytron 2, $985) with a sensing unit ($395). The multi-gas body can combine with specific sensor units to measure over 60 toxic gases including NH3. The expected life of an NH3 sensor is equal to or longer than 18 months. Twistik Transmitter Jiang and Sands (2000) measured NH3 concentrations in broiler buildings with an EC sensor (ETI series 4700 Twistik Transmitter), but did not give details about the sensor and its performance. (4) (3)

Solid State Sensor


The solid state or electronic NH3 sensor is a relatively new measurement method. It benefits from the boom of the electronic sensor technology. There are several types of these sensors that are sensitive to NH3 (Gpel and Schierbaum, 1991). There exist several advantages of solid state NH3 sensors, including simplicity, low price, quick response, and automatic measurement. Their limitations include low accuracy, drift, and

21

interference by humidity and other gases. Several types of NH3 sensors have been tested in animal houses. However, they were still in the development stage. Krause and Janssen (1990; 1991) first used a chemical NH3 sensor to measure NH3 distribution in animal houses. The sensor had a detection range of 11000 ppm, a response time <1 s and an accuracy of 10% between 4500 ppm. Krause (1993) described a test of a semiconductor NH3 sensor, but did not indicate whether it was the same sensor reported in 1990 and 1991. Berckmans et al. (1994) described a test of a solid state NH3 sensor, developed by the Interuniversity Micro Electronic Center (IMEC), Belgium, in livestock buildings. The sensor had a detection range of 0100 ppm NH3 and a response time of 1015 s. It was a thick film semiconducting metal oxide sensor consisting of a heater element, a dielectric layer, a contact layer, and a gas sensitive semi-conductive metal oxide layer (fig. 14). The conductivity of semiconducting metal oxide films at a certain temperature was influenced by the presence of NH3 gas in the surrounding atmosphere. The sensors optimum operating temperature was around 350400 C. Hess and Hgle (1994) tested an NH3 measuring system called SOLIDOX-NH3, manufactured in Germany, in animal houses for a few times but with difficulties.
Contact pad heater Dielectric layer Sensor contact Sensor material Sensor contact Dielectric layer

Contact pad heater 0 1 2 3 mm

Figure 14. Ammonia sensor developed by IMEC, Belgium.

Comparison Studies of Measuring Devices


When different devices are employed for NH3 concentration measurement at animal facilities, one question may naturally be raised: How comparable are they? So far, there have been several studies conducted to answer this question (table 4). Skewes and Harmon (1995) evaluated the AQT and the Gastec passive tubes against the Gastec active tubes in eight broiler houses. The authors concluded that the AQT estimated NH3 levels accurately at 2025 ppm and the passive tubes estimated average NH3 levels accurately at low levels of NH3. Their study indicated that the passive and the active tubes could not agree well, though both were products from the same company. Ni (unpublished) analyzed NH3 concentration data measured in the exhaust chimney in a pig house with a Drger EC sensor and with a CL analyzer. The correlation coefficient of NH3 concentrations was 0.55 from 3,900 paired data recorded by the two devices over more than one month.

22

Table 4. Summary of comparison studies of ammonia measuring devices. Compared devices 1. AQT 2. Gastec passive tubes 3. Gastec active tubes 1. Drger EC sensors 2. Kwik-Draw active tubes 3. Sensidyne passive tubes Test conditions and conclusions - Eight broiler houses: AQT and passive tubes agreed well with active tubes at 2025 ppm and lower levels, respectively. - Laboratory: all three devices agreed well. - Environmental chambers: EC sensors and active tubes differed significantly. - Layer houses: poor correlations existed among three devices. - Pig house: correlation coefficient r = 0.55 for 3,900 paired data points. - Pig house: test strips were precise and cost-effective as compared with Drger tubes. - Pig house: gas tubes could accurately measure NH3 using three sample averages. - Pig farm: agreement was reasonably good. Ref 1

1. Drger EC sensor 2. CL analyzer 1. pHydrion NH3 test strips 2. Drger tubes 1. Gas tubes 2. CL analyzer 1. Open-path FTIR 2. Open-path UV-DOAS

3 4 5 6

References: 1. Skewes and Harmon (1995); 3. Ni (Unpublished); 2. Wheeler et al. (2000b); 4. Dewey et al. (2000); 5. Parbst et al. (2000); 6, Secrest (2000).

Wheeler et al. (2000b) compared two Drger EC sensors, the Kwik-Draw gas detector tube, and the Sensidyne passive gas tube, in a lab, three environmental chambers, and three high-rise poultry houses. Although all devices agreed well in the lab test, there was a significant difference between the Drger sensor and the active tube in environmental chambers. There were poor correlations among the three types of sensors in poultry houses. Other studies showed satisfactory comparison results. Dewey et al. (2000) compared pHydrion NH3 test strips against Drger tubes, and concluded that the test strips provided a precise and cost-effective means of detecting NH3 concentrations in swine confinement buildings. Parbst et al. (2000) evaluated 0.25 to 3 ppm gas detection tubes (Drger Model 6733231) in finishing swine buildings in summer and winter against a CL analyzer (TEI Model 17C) being used in a long-term study (Heber et al. 2001). The authors concluded that NH3 concentrations could be accurately evaluated using the mean of three gas detection tube samples. Secrest (2000) co-located the optical paths of an open-path FTIR system and an open-path UVDOAS system. The FTIR path was 159 m and the UV-DOAS path was 150 m. The author concluded that the agreement was reasonably good (R2 = 0.94), although there was a 15 ppb difference between their responses at lower concentrations.

Calibration of Measuring Devices


Calibration of measuring devices assures data quality and provides information about characteristics of the devices, such as response, drift, linearity, stability, and precision.

23

Commercial measuring devices are usually calibrated by the supplier or manufacturer before shipping to the consumer following purchase, repair, or maintenance. However, although there are devices with claimed self-calibration capacity (Mount et al., 2001), calibration is required for most analytical instruments before initial use. Re-calibration after using the device for a certain time is necessary, because drift of the instrument occurs. Instrument manufacturers usually recommend calibration intervals, specify calibration procedure, and provide calibration accessories. Individual disposable measuring devices, such as active and passive gas detection tubes, cannot be calibrated by users, thus no quality control check is possible. However, these devices can be compared by randomly selecting tubes, e.g. 10%, from a batch of new tubes and comparing results against an acceptable reference device, e.g. a calibrated CL system. Calibration of an NH3 measuring device is usually realized by measuring NH3 concentration in certified gas mixtures of NH3 in nitrogen or NH3 in air. The difference between the known NH3 concentrations and the device outputs guides the correction of the measurement by adjusting system hardware or software, or by correcting concentration data during data processing. Only a small number of publications about NH3 studies in animal facilities indicate calibrations of measuring devices, which include NH3 analyzers (van 't Klooster and Heitlager, 1992; Heber et al., 2001), PAS gas analyzers (Rom, 1993), EC sensors (Wheeler et al., 1999), and UV DOAS and FTIR systems (Secrest, 2000). Ni and Heber (unpublished) evaluated three cylinders of 53.1, 33.2 and 9.33 ppm certified NH3 in air. The gases were ordered with the request that they be analyzed twice with at least one week apart between analyses with both analyses agreeing to within 1%. The gas company used FTIR as the analytical method. Upon receiving the gases, the NH3 directly from the cylinders and after 50% dilution using certified zero air was measured with an NH3 Analyzer (TEI Model 17C). The readings from the analyzer with and without dilution indicated good linearity of the analyzer (table 5). Table 5. Measurements of calibration gases with an ammonia analyzer when calibrated to the 53.1-ppm concentration. Certified NH3 concentrations ppm 53.1 33.2 9.33 29.6 (recertified) 9.6 (recertified) No dilution 53.1 40.5 5.77 16.9 6.0 Readings at NH3 analyzer, ppm Percentage difference 50% dilution 0.0% -18.0% +61.7% +75.1% +60.0% 26.6 20.0 2.80

However, when the analyzer was adjusted to correctly display the 53.1-ppm concentration, the measured concentrations was 40.5 and 5.77 ppm for the 33.2 and 9.33 ppm certified gases, respectively. The 53.1 ppm gas was confirmed to be the most reliable based on measurements with two other devices: NH3 detection tubes (Drger) and a photoacoustic IR NH3 analyzer (Chillgard IR Refrigerant Leak Detection System, MSA, Pittsburgh, PA). The two cylinders containing 33.2 and 9.33 ppm NH3 were therefore returned to the gas provider, who admitted the inaccuracy of the two gases, for recertification. However, the recertified gases still showed significant differences when using the 53.1-ppm NH3 as a calibration reference.

24

Discussion
Development and Application of Measurement Techniques
Measurement of NH3 concentrations at animal facilities began in the 1960s with wet chemistry (Valentine, 1964; Moum et al., 1969). With the development of sensor and analytical technology, more and more techniques became available for agricultural NH3 concentration measurement. So far, nine groups of techniques have been introduced and used at animal facilities. They were chosen to meet various objectives, different technical requirements, and budget constraints. There appears to be an increasing use of dry methods with direct readout in recent years. Some of these methods can handle larger number of measurements with immediate results, provide high frequency and high precision measurements, and enable computer-aided data logging and processing. Their successful application of these techniques advanced the characterization, quantification and modeling of agricultural NH3 emission significantly, e.g. by establishing NH3 emission factors for different sources and by revealing dynamic behaviors of NH3 release from animal wastes.

Data Quality of Agricultural Ammonia Measurement


The collection of high quality data is critical to any research program. Erroneous data are worse than no data because bad data mislead scientific conclusions, regulatory decisions, abatement technique evaluations, and health risk assessments. Information accompanying NH3 data reported in the literature is insufficient to provide an assessment of data quality. This is due to two problems. First, data quality has not received enough attention in past studies. In most published studies, there is little information about quality assurance and quality control (QAQC), e.g. calibration of measurement systems, assessment of precision and bias, etc. Second, proper and consistent data quality terminology is lacking, adding the difficulty in comparing research results. Different data quality indicators (DQIs) were used to describe the measurement quality in various reports. These DQIs include accuracy, inaccuracy, precision, error, sensitivity and standard deviation." This increased difficulties in easy comparison of research results. According to USEPA (1998), accuracy is a measure of the closeness of an individual measurement or the average of a number of measurements to the true value. Accuracy includes a combination of random error (precision) and systematic error (bias) components that result from sampling and analytical operations (fig. 15). The USEPA recommends using the terms precision and bias, rather than accuracy, to convey information usually associated with accuracy. Precision is a measure of agreement among replicate measurements of the same property, under prescribed similar conditions. Bias is the systematic or persistent distortion of a measurement process that causes errors in one direction. To improve data quality in future studies and investigations, errors in agricultural NH3 measurement and measures to reduce these errors need more attention by researchers who voluntarily following USEPA guidelines for QAQC if they are not required to do so.

Sources of Measurement Errors


When conducting NH3 measurement at animal facilities, there are several potential sources of errors that may result from sampling, measurement, and data processing (table 6). While most

25

of these error sources are common in any scientific research program, a few of them deserve particular attention.

* * * * * * * *

* *

* * *

* *

High bias + low precision = low accuracy

Low bias + low precision = low accuracy

* ** ** * ** High bias + high precision = low accuracy

** * ** * **

Low bias + high precision = high accuracy

Figure 15. Random measurement uncertainties and measurement bias: shots at a target. Source: USEPA (1998). Calibration Gas Ammonia measuring devices used in agricultural are usually field-calibrated under steady-state conditions by applying a step input using NH3 calibration gas and a zero input using zero air. Calibration gas is a serious potential source of systematic error in the NH3 concentration measurement. This was demonstrated by the study of Ni and Heber (unpublished) and explained in table 5, which showed obvious disagreement among NH3 gases with certified concentrations. This case raised a critical question about the reliability of calibration gases. If a measuring system is calibrated against an incorrect NH3 concentration, the systematic error passes on to collected and processed data. The cause of this error goes back to NH3 gas standards, and is related to calibration gas manufacturers: the analytical method they use and the QAQC they follow. The National Institute of Standards and Technology (NIST) provides U.S. industry, government, and the public with measurements, standards, and information services. The USEPA traceability protocol for assay and certification of gaseous calibration standards (USEPA, 1997) allows for Gas Manufacturers Intermediate Standards (GMIS) and NIST Traceable Reference Materials (NTRM) Gas Calibration Standards to be prepared by gas manufacturers and certified by NIST. Unlike some other gases with agricultural origin (e.g. hydrogen sulfide (H2S), methane (CH4), carbon dioxide (CO2), and sulfur dioxide (SO2)), NH3 is not yet available as a compressed gas with Standard Reference Materials from NIST (USEPA, 1997). High Frequency Variation of Ammonia Concentrations Ammonia concentrations at animal facilities are almost always under transient conditions with high frequency and large magnitude of variations (Ni et al., 2000b). These concentration

26

variations do not introduce serious measurement errors to sensors that provide TWA concentrations (e.g. passive gas tubes). Table 6. Some potential error sources in ammonia concentration measurement. Processes Sampling Sources of errors Diurnal gas concentration variation Comments Frequently repeated measurements covering day and night may reduce error Long-period measurements covering warm & cold seasons may reduce error Carefully selected multi-location sampling may reduce error Affects response time

Seasonal gas concentration variation Spatial gas concentration variation Gas adsorption/desorption from contacting surfaces of sampling system Condensation in sampling system Dust in sampling system Leaks in sampling system Sample volume measurement Disturbance of NH3 source Operation of sampling Measurement Calibration gas Calibration procedure Instrument sensitivity and precision Instrument stability Inadequate frequency response Interferences of water vapor and other gases Operation of measurement Data processing Temperature data Atmospheric pressure data Data round off Gas equilibrium time selection

Absorbs NH3, affects sensor operation Use filters Careful test and system improvement may prevent leaks For example when using wet chemistry When using sampling chambers Human errors Potentially serious systematic error QAQC practices may reduce error Expected random error Random or systematic errors Random error Develop and use new calibration methodology and select better devices Human errors For converting mass concentrations For converting mass concentrations Expected random error Applies only to multiple-loation sampling

27

However, they do introduce errors to instruments (e.g. FTIR, PAS, and NOx analyzer) that provide real-time NH3 concentrations during measurement but have inadequate frequency responses compared with the dynamic changes in concentrations. Accurate dynamic measurement requires a small time constant of the instrument (Doebelin, 1983). The magnitude and correction of this type of error and its relationship with specific measuring device types, sampling method, animal facility size, etc. need more investigation. Furthermore, the true frequencies of NH3 concentrations under various conditions (e.g. in open air, inside buildings, with different sizes of buildings, at different seasons, with different weather, etc.) have not yet been fully characterized and deserve future study. Interference of Water Vapor, Other Gases and Particular Matter Except for the pH paper method, almost all of the NH3 measurement techniques were initially developed for non-agricultural use. They were introduced to the animal industry after successful use in other fields. For instance, the Chemcassette Monitor that has existed for over 30 years was first reported in 2000 for agricultural use. The conditions under which the instruments were developed and factory-calibrated are very different from the conditions at animal facilities. Air produced from animal facilities is a mixture of a large number of gases plus relatively high moisture content compared with the commercial calibration gases. Some of the NH3 measuring devices are sensitive to water vapor and gases other than NH3. Interferences are therefore possible during field measurements. Three of the six comparison studies summarized in table 4 demonstrated at least some degree of disagreement between the tested measuring devices. The increasing inconsistency of the three devices between the lab, environmental chamber, and commercial layer house reported by Wheeler et al. (2000b) could probably be explained by increasing interferences among the three test locations. The gases that interfere with detection tubes may be more prevalent in commercial poultry houses than one might expect (Wheeler et al., 2000b). No discussion about development or improvement of NH3 measuring devices specifically targeting the agricultural environment was found in the literature. The potential errors caused by agricultural air interference are not fully understood and are not compensated in field data. Therefore, there is an urgent need to develop test methodology and conduct tests in order to determine field performance of available measuring devices. Recommendations on the use of these devices and the field data they produced will help to improve agricultural NH3 measurement significantly.

Selection of Sampling Techniques


Sampling location and time is critical to obtain high quality data. Different sampling locations may result in wide variations in measurement data because of spatial NH3 differences. Measurements of varying concentrations that cover excessively short periods produce data with serious temporal limitations. However, measurement objectives play an important role in selecting sampling location and time. For example, for animal or human exposure studies, sampling locations should be in animal or human respiration zones, whereas the best sampling locations are the building air exhausts for emission measurement. In the mechanically-ventilated negative-pressure animal house, the sampling position can be chosen at the exhaust fans for emission study. The advantage of this technique is that the gas concentration in the exhaust represents the outgoing gas concentration. Since the ventilation rate can also be measured in the exhaust(s), it is favorable for obtaining relatively accurate gas

28

emission data. This sampling technique was reported by Berckmans and Ni (1993), Hartung et al. (1997), and Heber et al. (2001). The temporal variations of NH3 concentrations shown in fig. 3 demonstrated that it is important to select proper sampling time. Sampling time should be arranged to cover peak and valley concentrations during the day, especially when there are significant temperature and airflow rate fluctuations, to obtain daily mean concentration, whether short duration sampling (e.g. active gas detection tube) or long-duration sampling (e.g. passive gas tube or wet chemistry) techniques are used. Based on the same principle, sampling should be designed to cover the low concentration season (usually summer) and high concentration season (usually winter) if an annual mean concentration is to be obtained. It is clear that low frequency sampling results in poor representation of the true NH3 fluctuation pattern and therefore unreliable mean NH3 concentrations. The higher the sampling frequency is, the better the data resolution, and the more accurate the mean value. According to the Nyquist Theorem, the sampling frequency should be at least twice the maximum frequency of the signal that is being sampled (Finkelstein and Grattan, 1994). Of course, high frequency sampling is subject to some technical restrictions like the capacity and response time of the measuring device.

Selection of Measurement Techniques


Selection of measurement techniques should be based on research objectives, coupled with the existing capabilities of the research institution and the budget constraints of the research. Cost of the techniques is one of the most important factors to be considered in almost all research projects. Thus, capital and operating costs may need to be assessed with the performance of the technique. A single-use sensor can only provide one measurement although it is relatively inexpensive. If large number of measurement data is required, the cost per measurement using a high-priced instrument with multi-use sensors may be less expensive than using low-price single-use sensors. In many cases, small numbers of short-term samples cannot satisfy the accuracy requirements of a careful field investigation. Techniques that produce large quantities of data should therefore be considered. Some expensive instruments, like IR analyzers and NH3 analyzers, are usually only used at institutions conducting intensive research on agricultural NH3. Standard wet chemistry requires analytical instruments that may already exist at many institutions. When the cost of analytical instruments does not need to be considered, wet chemistry methods are inexpensive and affordable techniques. They are especially useful with small sample numbers. The pH-paper-based test kits are appropriate for obtaining mere indications of in-building NH3 concentrations when accuracy is not a priority. Applications of high and low sensitivity measuring devices are generally related to indoor and outdoor NH3 measurements, respectively. Indoor NH3 concentrations are usually above 1.0 ppm, and almost all the techniques reviewed in this paper are compatible. Since outdoor NH3 may be at ppb levels, some sensors, e.g. gas detection tubes, are not appropriate due to lack of sensitivity. Measuring devices with short response times (e.g. less than 2 min) are required to properly study the dynamic behavior and diurnal variations of NH3 concentrations. Sensors with long response times, e.g. passive gas tubes, are very good when only TWA data are needed. To date, however, it is still unknown which techniques provide results that are the closest to the true NH3 concentrations under agricultural field conditions. The tests shown in table 4 only

29

compared selected techniques with unknown characteristics against each other, not against a standard technique, which does not yet exist. Therefore, a standard technique and relevant methodologies need to be developed, and existing NH3 concentration measuring devices need to be tested and compared.

Standards for Agricultural Ammonia Measurement


Because methodologies that are proven to be scientifically sound are still unavailable, many of the field studies have not produced reliable data. Some European researchers believe that about 80% of the publications of agricultural NH3 emissions are not useable to establish annual emission factors (Gallmann and Hartung, 2000). This has created a lack of confidence in reported data and has made comparison of research results difficult or impossible. The present situation leads to the conclusion that there is an urgent need to develop standards for agricultural NH3 measurement and relevant methodologies in this specific field. The following standards are needed: Technical terms: Standardize technical terms for this specific research field including those for describing processes (e.g. NH3 production, generation, volatilization, release, emission, etc.), presenting data (e.g. concentration, emission flux, emission rate, emission per animal unit, emission per animal place, annual emission factor, etc.), and indicating data qualities (e.g. precision, bias, accuracy, etc.). Calibration gases: Define approved analytical instruments and procedures used by calibration gas providers for certifying gases. This standard will guide researchers on selecting gas providers and will assure quality of certifications. Sampling devices and procedures: Standardize closed and open sampling devices (e.g. dimensions and structures of sampling chambers, configuration of micro-metrological sampling, materials used in sampling systems, etc.), and sampling procedures (e.g. sampling location, interval, frequency, duration, season, etc.). Measuring devices: Standardize performance requirements (e.g. precision, sensitivity, response time, etc.) of measuring devices (including those for NH3 concentration and airflow rate measurements) for various measurement objectives (e.g. compliance, emission factor determination, abatement evaluation and health risk assessment, etc.)

Conclusions
1. Temporal and spatial variations of NH3 concentrations at animal facilities are major technical difficulties for NH3 sampling and measurement. 2. More and more NH3 measuring devices have been used for studying NH3 concentrations in agricultural environments, providing more technical possibilities for research with various objectives and advancing the studies of NH3 emissions and tests of abatement techniques. 3. Some measuring devices exhibit different levels of performance under different measurement conditions. More investigations are needed to assess their performance under field conditions in agriculture. 4. Due to lack of scientifically sound methodologies and insufficient QAQC for NH3 measurement, there is a lack of confidence in the quality of reported data. Comparison of research results is difficult or impossible in many cases. 5. Some sources of errors associated with sampling and measurement need careful study.

30

6. There is a potentially serious error in certifications of NH3 calibration gases due to lack of NIST standards for analyzing NH3. 7. Methodologies for assessing sampling and measuring devices and procedures for agricultural NH3 studies need to be developed. 8. Development of standards of technical terms, calibration gases, sampling devices and procedures, and measuring devices is an urgent need. 9. Cooperation of instrument developers/manufacturers, calibration gas providers, regulators, standard agencies, and agricultural research scientists is necessary to improve agricultural NH3 measurement.

Acknowledgements
This study was supported, in part, by the Purdue Center for Animal Waste Management Technologies funded by the USEPA, and the Purdue University Agricultural Research Program. Part of this paper was based on some previous work, which was partially supported by Hangzhou Rural Energy Office, China, and Catholic University of Leuven, Belgium.

References
Aarnink, A.J.A., M.J.M. Wagemans, and A. Keen. 1993. Factors affecting ammonia emission from housing for weaned piglets. In Proceedings of the First International Symposium on Nitrogen Flow in Pig Production and Environmental Consequences, eds, M.W.A. Verstegen et al. 286-294. Wageningen, The Netherlands, 8-11 June. Wageningen: Pudoc Scientific Publishers. Amon, B., J. Boxberger, T. Amon, A. Gronauer, G. Depta, S. Neser, and K. Schfer. 1997. Methods for measuring emissions from agrarian sources: FTIR measurement techniques with White-Cell, large chamber or open-path. In International Symposium on Ammonia and Odour Control from Animal Production Facilities, eds, J.A.M. Voermans and G.J. Monteny. Vol. I. 161-167. Vinkeloord, The Netherlands, 6-10 Oct. Rosmalen, The Netherlands: NVTL. Andersson, M. 1995. The effect of different manuring systems on ammonia emissisns in pig buildings. Institutionen fr Jordbrukets Biosystem och Teknologi, Sveriges Lantbruksuniversitet, Lund, Sweden. Report No. 100. 41 p. Aneja, V.P., J.P. Chauhan, and J.T. Walker. 2000. Characterization of atmospheric ammonia emissions from swine waste storage and treatment lagoons. Journal of Geophysical Research 105:11535-11545. Aneja, V.P., E.P. Stahel, H.H. Rogers, A.M. Witherspoon, and W.W. Heck. 1978. Calibration and performance of a thermal converter in continuous atmospheric monitoring of ammonia. Analytical Chemistry 50(12):1705-1707. Asteraki, E.J., R.A. Matthews, and B.F. Pain. 1997. Ammonia emissions from beef cattle bedded on straw. In International Symposium on Ammonia and Odour Control from Animal Production Facilities, eds, J.A.M. Voermans and G.J. Monteny. Vol. I. 343-347. Vinkeloord, The Netherlands, 6-10 Oct. Rosmalen, The Netherlands: NVTL. Berckmans, D. and J.Q. Ni. 1993. Field test installation for on-line continuous measurement of total ammonia emission from animal houses. In Proceedings of the International Conference for Agricultural Machinery & Process Engineering. Vol. 2. 393-402. Seoul, Korea, 19-22 October.

31

Berckmans, D., J.Q. Ni, V. Goedseels, J. Roggen, G. Huyberechts, and R. Van Overstraeten. 1994. Feasibility study of the use of an existing sensor for the measurement and control of ammonia emission from animal production unit. K.U.Leuven and IMEC, Leuven, Belgium. VLIM/H/9032, Final report. 57 p. Bicudo, J.R., C.L. Tengman, L.D. Jacobson, and J.E. Sullivan. 2000. Odor, hydrogen sulfide and ammonia emissions from swine farms in Minnesota. In Conference Proceedings Odors and VOC Emissions 2000, Session 8. 20. Cincinnati, Ohio, USA, 16-19 April. Alexandria, VA : Water Environment Federation. Brewer, S.K. and T.A. Costello. 1997. In-situ measurement of ammonia volatilization from broiler litter using an enclosed air chamber. ASAE Meeting Paper No. 97-4068. St. Joseph, Mich.: ASAE. Brunsch, R. 1997. Methodical aspects relating the results of multigas monitoring. In International Symposium on Ammonia and Odour Control from Animal Production Facilities, eds, J.A.M. Voermans and G.J. Monteny. Vol. I. 185-191. Vinkeloord, The Netherlands, 6-10 Oct. Rosmalen, The Netherlands: NVTL. Busse, F.-W. 1993. Comparison measurements of the house climate in swine stables with and without respiratory diseases or cannibalism. In Livestock Environment. Fourth International Symposium, eds, E. Collins and C. Boon. 904-908. University of Warwick, Coventry, England, 6-9 July. American Society of Agricultural Engineers, Michigan, USA. Carr, L.E., F.W. Wheaton, and L.W. Douglass. 1990. Empirical models to determine ammonia concentrations from broiler chicken litter. Transactions of the ASAE 33(4):1337-1342. Choinire, Y., B. Marquis, and G. Gingas. 1997. Ammonia and contaminant concentrations with conventional versus pit ventilation in finishing pig units. In International Symposium on Ammonia and Odour Control from Animal Production Facilities, eds, J.A.M. Voermans and G.J. Monteny. Vol. I. 365-372. Vinkeloord, The Netherlands, 6-10 Oct. Rosmalen, The Netherlands: NVTL. Crook, B., J.F. Robertson, Glass Sat, E.M. Botheroyd, J. Lacey, and M.D. Topping. 1991. Airborne dust, ammonia, microorganisms, and antigens in pig confinement houses and the respiratory health of exposed farm-workers. American Industrial Hygiene Association Journal 52(7):271-279. Curtis, S.E., C.R. Anderson, J. Simon, A.H. Jensen, D.L. Day, and K.W. Kelley. 1975. Effects of aerial ammonia, hydrogen sulfide and swine-house dust on the rate of gain and respiratory-tract structure in swine. Journal of Animal Science 41:735-739. De Praetere, K. and W. van Der Biest. 1990. Airflow patterns in piggeries with fully slatted floors and their effect on ammonia distribution. Journal of Agricultural Engineering Research 46:31-44. Demmers, T.G.M., L.R. Burgess, Short, J L, V.R. Phillips, J.A. Clark, and C.M. Wathes. 1999. Ammonia emissions from two mechanically ventilated UK livestock buildings. Atmospheric Environment 33(2):217-227. den Brok, G.M. and N. Verdoes. 1997. Slurry cooling to reduce ammonia emission from pig houses. In International Symposium on Ammonia and Odour Control from Animal Production Facilities, eds, J.A.M. Voermans and G.J. Monteny. Vol. II. 441-447. Vinkeloord, The Netherlands, 6-10 Oct.. Rosmalen, The Netherlands: NVTL. Dewey, C.E., B. Cox, and J. Leyenaar. 2000. Measuring ammonia concentrations in the barn using the Draeger TM and pHydrion TM tests. Swine Health and Production 8(3):127-131. Doebelin, E.O. 1983. Measurement Systems : Application and Design. 3 ed. New York: McGraw-Hill. 867 p.

32

Donham, K.J., S.J. Reynolds, P. Whitten, J.A. Merchant, L. Burmeister, and W.J. Popendorf . 1995. Respiratory dysfunction in swine production facility workers -dose-response relationships of environmental exposures and pulmonary-function. American Journal of Industrial Medicine 27(3):405-418. Elwinger, K. and L. Svensson. 1996. Effect of dietary protein content, litter and drinker type on ammonia emission from broiler houses. Journal of Agricultural Engineering Research 64:197-208. Elzing, A. and D. Swierstra. 1993. Ammonia emission measurements in a model system of a pig house. In Proceedings of the First International Symposium on Nitrogen Flow in Pig Production and Environmental Consequences, eds, M.W.A. Verstegen et al. 280-285. Wageningen, The Netherlands, 8-11 June. Wageningen, The Netherlands: Pudoc Scientific Publishers. Ferguson, N.S., R.S. Gates, A.H. Cantor, J.L. Taraba, A.J. Pescatore, M.L. Straw, M.J. Ford, L.W. Turner, and D.J. Burnham. 1997. Effects of dietary crude protein on growth, ammonia concentration and litter composition in broilers. ASAE Meeting Paper No. 974069. St. Joseph, Mich.: ASAE. Finkelstein, L. and K.T.V. Grattan. eds. 1994. Concise encyclopedia of measurement & instrumentation. Advances in systems, control, and information engineering. Oxford: Pergamon Press. 434 p. Gallmann, E. and E. Hartung. 2000. Evaluation of the emission rates of ammonia and greenhouse gases from swine housings. In Proc. 2nd International Conference on Air Pollution from Agricultural Operations. 92-99. Des Moines, IA, 9-11 Oct. St. Joseph, Michigan: ASAE. Gpel, W. and K.-D. Schierbaum. 1991. Chemical and Biochemical Sensors, Part 1. In Sensors, A Comprehensive Survey, eds. W. Gpel et al. pp. 119-158. Weinheim: VCH. Groenestein, C.M., H.M. Vermeer, and J.M.G. Hol. 1997. Ammonia emission and feedinginduced activity from houses with sows kept individually and in groups. In International Symposium on Ammonia and Odour Control from Animal Production Facilities, eds, J.A.M. Voermans and G.J. Monteny. Vol. II. 553-560. Vinkeloord, The Netherlands, 6-10 Oct. Rosmalen, The Netherlands: NVTL. Harris, D.B., R.C. Shores, and L.G. Jones. 2001. Ammonia emission factors from swine finishing operations. In 10th Annual International Emission Inventory Conference, One Atmosphere, One Inventory, Many Challenges. Adams Mark Hotel, Denver, Colorado, May 1-3. Washington D.C.: USEPA. Hartung, E., T. Jungbluth, and W. Bscher. 1997. Reduction of ammonia and odor emissions from a piggery with biofilters. ASAE Meeting Paper No. 97-4126. St. Joseph, Mich.: ASAE. Hartung, E., T. Jungbluth, and W. Bscher. 2001. Reduction of ammonia and odor emissions from a piggery with biofilters. Transactions of the ASAE 44(1):113-118. Hashimoto, A.G. 1972. Aeration under caged lying hens. Transactions of the ASAE 15(6):11191123. Hauser, R.H. and D.W. Flsch. 1993. The quality of poultry-house in alternative systems for farming laying hens. In Fourth International Symposium of Livestock Environment, eds, E. Collins and C. Boon. 671-677. University of Warwick, Coventry, England, 6-9 July. Michigan, USA: ASAE. Heber, A.J., R.K. Duggirala, J.Q. Ni, M.L. Spence, B.L. Haymore, V.I. Adamchuk, D.S. Bundy, A.L. Sutton, and K.M. Keener. 1997a. Field performance of a pit additive tested in

33

commercial grow-finish houses. In Swine Day 97. 97-106. August 28. Purdue University Cooperative Extension Service, West Lafayette, Indiana. Heber, A.J., R.K. Duggirala, J.Q. Ni, M.L. Spence, B.L. Haymore, V.I. Adamchuk, D.S. Bundy, A.L. Sutton, D.T. Kelly, and K.M. Keener. 1997b. Manure treatment to reduce gas emissions from large swine houses. In International Symposium on Ammonia and Odour Control from Animal Production Facilities, eds, J.A.M. Voermans and G.J. Monteny. Vol. II. 449-458. Vinkeloord, The Netherlands, 6-10 Oct. Rosmalen, The Netherlands: NVTL. Heber, A.J., T.T. Lim, J.Q. Ni, and A.L. Sutton. 2000a. Odor and gas emission from anaerobic treatment of swine manure. Final report to Indiana's Value-Added Agricultural Grant Program, Office of the Commissioner of Agriculture. Purdue University, West Lafayette, Indiana. December. 47 p. Heber, A.J., J.Q. Ni, T.T. Lim, C.A. Diehl, A.L. Sutton, R.K. Duggirala, B.L. Haymore, D.T. Kelly, and V.I. Adamchuk. 2000b. Effect of a manure additive on ammonia emission from swine finishing buildings. Transactions of the ASAE 43(6):1895-1902. Heber, A.J., J.Q. Ni, B.L. Haymore, R.K. Duggirala, and K.M. Keener. 2001. Air quality and emission measurement methodology at swine finishing buildings. Transactions of the ASAE (in press). Heinrichs, P. and J. Oldenburg. 1993. Effect of protein feeding on gaseous ammonia emissions and slurry loading with nitrogen in fattening pigs. In Proceedings of the First International Symposium on Nitrogen Flow in Pig Production and Environmental Consequences, eds, M.W.A. Verstegen et al. 336-339. Wageningen, The Netherlands, 8-11 June 1993. Pudoc Scientific Publishers, Wageningen. Hendriks, J.G.L. and M.G.M. Vrielink. 1997. Reducing ammonia emission from pig houses by adding or producing organic acids in pig slurry. In International Symposium on Ammonia and Odour Control from Animal Production Facilities, eds, J.A.M. Voermans and G.J. Monteny. Vol. II. 493-451. Vinkeloord, The Netherlands, 6-10 Oct. Rosmalen, The Netherlands: NVTL. Hess, H.J. and T. Hgle. 1994. Comparative measuring on ammonia emissions. Landtechnik 49(6):362-363. Hesse, D. 1994. Comparison of different old and new fattening pig husbandrys with focus on environment and animal welfare. In Proceedings of the XII World Congress on Agricultural Engineering. 559-566. Milano, 29 Aug. 1 Sept. Hoy, S. 1995. Studies on the use of multi-gas monitoring in animal houses. Tierarztliche Umschau 50:115-123. Hoy, S., K. Mller, and R. Willig. 1994. Ammonia concentrations in pig houses with different types of floors -- deep-litter husbandry with bioactivator compared to fully slatted floor husbandry (Zur Ammoniak-Konzentration bei zwei Systemen der Mastschweinehaltung auf Tiefstreu mit Bioaktivator im Vergleich zur Vollspaltenbodenhaltung). Tierrztliche Umschau 49(7):414-416, 418-420. Hoy, S. and R. Willig. 1994. Results of continuous measurement of ammonia in the air of pig houses, using electrochemical sensors [Ergebnisse von Verlaufsmessungen der Ammoniak-Konzentration in Schweinestallen mit Hilfe der Sensortechnik]. Monatshefte fr Veterinrmedizin 49:37-41. Hoy, S., R. Willig, and I. Buchholz. 1992. Results from continuous measurements of ammonia in keeping fattening pigs on deep litter with additives in comparison with housing on slatted floors. In Workshops: Deep Litter Systems for Pigs, ed. J.A.M. Voermans. 37-50. Rosmalen, the Netherlands, September 21-22.

34

Jacobson, L.D., B. Hetchler, K.A. Janni, and L.J. Johnston. 1998. Odor and gas reduction from sprinkling soybean oil in a pig nursery. ASAE Meeting Paper No. 984125. St. Joseph, Mich.: ASAE. Jacobson, L.D., K.A. Janni, P.E. Arellano, and C.J. Pijoan. 1992. Winter swine ventilation evaluation using air quality criteria. ASAE Meeting Paper No. 924039. St. Joseph, Mich.: ASAE. Jiang, J.K. and J.R. Sands. 2000. Odour and ammonia emission from broiler farms. Rural Industries Research and Development Corporation, Kingston, Australia. Publication No. 00/2. February. 94 p. Johnston, N.L., C.L. Quarles, D.J. Fagerberg, and D.D. Caveny. 1981. Evaluation of yucca saponin on broiler performance and ammonia suppression. Poultry Science 60:22892292. Johnston, S. 1991. Fourier transform infrared: a constantly evolving technology. New York: E. Horwood. 340 p. Jungbluth, T., G. Brose, and E. Hartung. 1997. Ammonia and greenhouse gas emissions from dairy barns. ASAE Meeting Paper No. 97-4127. St. Joseph, Mich.: ASAE. Jungbluth, T. and W. Bscher. 1996. Reduction of ammonia emissions from piggeries. ASAE Meeting Paper No. 96-4091. St. Joseph, Mich.: ASAE. Kamin, H., J.C. Barber, S.I. Brown, C.C. Delwiche, D. Grosjean, J.M. Hales, J.L.W. Knapp, E.R. Lemon, C.S. Martens, A.H. Niden, R.P. Wilson, and J.A. Frazier. 1979. Ammonia. Baltimore: University Park Press. 384 p. Kant, P.P.H., M.C. Verboon, and Huis in 't Veld. 1992. Ammonia emission measurements with the Lindvall box - A report on the measurements carried out at Waiboerhoeve Experimental Farm in 1989-1991 (Ammoniak-emissiemetingen met de Lindvalldoos Inventarisatie van de metingen op de Waiboerhoeve in 1989-1991). Lelystad en IMAG, Wageningen. 48 p. Kay, R.M. and P.A. Lee. 1997. Ammonia emission from pig buildings and characteristics of slurry produced by pigs offered low crude protein diets. In International Symposium on Ammonia and Odour Control from Animal Production Facilities, eds, J.A.M. Voermans and G.J. Monteny. Vol. I. 253-260. Vinkeloord, The Netherlands, 6-10 Oct. Rosmalen, The Netherlands: NVTL. Kay, R.M., B.F. Pain, C.R. Clarkson, and T.H. Misselbrook. 1992. The effect of a slurry additive on odour and ammonia emissions from pig buildings. Animal Production 54(3):2p-22. Keck, M., W. Bscher, and T. Jungbluth. 1994. Influence on the ammonia emission by under and overfloor extraction in a pig house. In AgEng '94. International Conference on Agricultural Engineering. 230-231. Milano, 29 Aug. - 1 Sept. Krause, K.-H. 1993. On dispersion of ammonia emissions (Zur Ausbreitung van Ammoniakemissionen). Landtechnik 48:562-569. Krause, K.-H. and J. Janssen. 1990. Measuring and simulation of the distribution of ammonia in animal houses. In Room Vent '90, Session C-7. 1-12. Krause, K.-H. and J. Janssen. 1991. Modelling the dispersion of ammonia within animal houses. In Cost 681 Expert Odours Group Workshop: Odour and Ammonia Emissions from Livestock Farming, eds, V.C. Nielsen et al. 71-80. Silsoe, UK, 26-29 March 1990. Elsevier Applied Science, London/New York. Krieger, R., J. Hartung, and A. Pfeiffer. 1993. Experiments with a feed additive to reduce ammonia emissions from pig fattening housing - preliminary results. In Proceedings of the First International Symposium on Nitrogen Flow in Pig Production and Environmental

35

Consequences, eds, M.W.A. Verstegen et al. 295-300. Wageningen, 8-11 June. Pudoc Scientific Publishers, Wageningen. Kroodsma, W., J.W.H. HuisintVeld, and R. Scholtens. 1993. Ammonia emission and its reduction from cubicle houses by flushing. Livestock Production Science 35:293-302. Kurvits, T. and T. Marta. 1998. Agricultural NH3 and NOx emissions in Canada. Environmental Pollution 102(Suppl. 1):187-194. Leichnitz, K. 1985. Analyses de l'air et analyse technique des gaz l'aide des tubes ractifs Drger. 6 ed. 286 p. Leithe, W. 1971. The analysis of air pollutants. Ann Arbor, USA: Ann Arbor Science Publishers. Lindvall, T., O. Norn, and L. Thyselius. 1974. Odor reduction for liquid manure systems. Transactions of the ASAE 17:510-512. Liu, Q., D.S. Bundy, and S.J. Hoff. 1993. Utilizing ammonia concentrations as an odor threshold indicator for swine facilities. In Livestock Environment. Fourth International Symposium, eds, E. Collins and C. Boon. 678-685. Coventry, University of Warwick, England, 6-9 July. ASAE, St. Joseph, Michigan. Maghirang, R.G. and H.B. Manbeck. 1993. Dust, ammonia, and carbon dioxide emissions from a poultry house. ASAE Meeting Paper No. 934056. St. Joseph, Mich.: ASAE. Maghirang, R.G., H.B. Manbeck, W.B. Roush, and F.V. Muir. 1991. Air contaminant distributions in a commercial laying house. Transactions of the ASAE 34:2171-2180. Mannebeck, H. and J. Oldenburg. 1991. Comparison of the effects of different systems on ammonia emissions. In Cost 681 Expert Odours Group Workshop: Odour and Ammonia Emissions from Livestock Farming, eds, V.C. Nielsen et al. 42-49. Silsoe, UK, 26-29 March 1990. Elsevier Applied Science, London. McGinn, S.M. and H.H. Janzen. 1998. Ammonia sources in agriculture and their measurement. Canadian Journal of Soil Science 78:139-148. Meyer, V. and D. Bundy. 1991. Farrowing building air quality survey. ASAE Meeting Paper No. 914012. St. Joseph, Mich.: ASAE. Misselbrook, T.H., B.F. Pain, and D.M. Headon. 1998. Estimates of ammonia emission from dairy cow collecting yards. Journal of Agricultural Engineering Research 71(2):127-135. Monteny, G.J. and J.W. Erisman. 1998. Ammonia emission from dairy cow buildings: A review of measurement techniques, influencing factors and possibilities for reduction. Netherlands Journal of Agricultural Science 46(3-4):225-247. Moum, S.G., W. Seltzer, and T.M. Goldhaft. 1969. A simple method of determining concentrations of ammonia in animal quarters. Poultry Science 48:347-348. Mount, G.H., B.P. Rumburg, B.L. Lamb, J.R. Havig, H. Westberg, K.A. Johnson, and R.L. Kincaid. 2001. DOAS measurement of atmospheric ammonia emissions at a dairy. In 10th Annual International Emission Inventory Conference, One Atmosphere, One Inventory, Many Challenges. Adams Mark Hotel, Denver, Colorado, 13 May. Washington D.C.: USEPA. Myers, J., T. Kelly, C. Lawrie, and K. Riggs. 2000. Environmental technology verification report: Opsis Inc. AR-500 Ultraviolet Open-Path Monitor. Battelle, Columbus, OH. September. 31 p. Neser, S., G. Depta, B. Stegbauer, A. Gronauer, and H. Schn. 1997. Mass balance of the compounds nitrogen and carbon in housing systems for laying hens. In International Symposium on Ammonia and Odour Control from Animal Production Facilities, eds, J.A.M. Voermans and G.J. Monteny. Vol. I. 129-136. Vinkeloord, The Netherlands, 6-10 Oct. Rosmalen, The Netherlands: NVTL.

36

Ni, J.Q. 1998. Emission of carbon dioxide and ammonia from mechanically ventilated pig house. Ph.D. Thesis. Department of Agricultural Engineering, Catholic University of Leuven. Leuven, Belgium. 227 p. Ni, J.Q., A.J. Heber, C.A. Diehl, and T.T. Lim. 2000a. Ammonia, hydrogen sulphide and carbon dioxide from pig manure in under-floor deep pits. Journal of Agricultural Engineering Research 77(1):53-66. Ni, J.Q., A.J. Heber, T.T. Lim, C.A. Diehl, R.K. Duggirala, B.L. Haymore, and A.L. Sutton. 2000b. Ammonia emission from a large mechanically-ventilated swine building during warm weather. Journal of Environmental Quality 29(3):751-758. Nicks, B., A. Dsiron, and B. Canart. 1997. Deep litter materials and the ammonia emissions in fatterning pig houses. In International Symposium on Ammonia and Odour Control from Animal Production Facilities, eds, J.A.M. Voermans and G.J. Monteny. Vol. I. 335-342. Vinkeloord, The Netherlands, 6-10 Oct. Rosmalen, The Netherlands: NVTL. Nicks, B., D. Marlier, and B. Canart. 1993. Air pollution levels in pig houses. In Fourth International Symposium of Livestock Environment, eds, E. Collins and C. Boon. 635642. University of Warwick, Coventry, England, 6-9 July. American Society of Agricultural Engineers, Michigan, USA. Osada, T., H.B. Rom, and P. Dahl. 1998. Continuous measurement of nitrous oxide and methane emission in pig units by infrared photoacoustic detection. Transactions of the ASAE 41(4):1109-1114. Parbst, K.E., K.M. Keener, A.J. Heber, and J.Q. Ni. 2000. Comparison of a low cost and high cost odor monitoring equipment in a commercial swine finishing house. Applied Engineering in Agriculture 16(6):693-699. Patni, N.K. and S.P. Clarke. 1991. Transient hazardous conditions in animal buildings due to manure gas released during slurry mixing. Applied Engineering in Agriculture 7(4):478484. Pfeiffer, A., F. Arends, H.J. Langholz, and G. Steffens. 1993. The influence of various pig housing systems and dietary protein levels on the amount of ammonia emissions in the case of fattening pigs. In Proceedings of the First International Symposium on Nitrogen Flow in Pig Production and Environmental Consequences, eds, M.W.A. Verstegen et al. 313-317. Wageningen (Doorwerth), The Netherlands, 8-11 June. Pudoc Scientific Publishers, Wageningen. Phillips, V.R., S.J. Lane, and L.R. Burgess. 2000. A technique for measuring ammonia emissions from the individual parts of a livestock building. In Air Pollution from Agricultural Operations, Proceedings of the Second International Conference. 84-91. Des Moines, Iowa, 9-11 October. ASAE, Michigan. Phillips, V.R., D.S. Lee, R. Scholtens, J.A. Garland, and R.W. Sneath. 2001. A review of methods for measuring emission rates of ammonia from livestock buildings and slurry or manure stores, part 2: monitoring flux rates, concentrations and airflow rates. Journal of Agricultural Engineering Research 78(1):1-14. Pranitis, D. and M. Meyerhoff. 1987. Continuous monitoring of ambient ammonia with a membrane-electrode-based detector. Analytical Chemistry 59(October 1):2345-2350. Pratt, S.E., L.M. Lawrence, T. Barnes, D. Powell, and L.K. Warren. 2000. Measurement of ammonia concentrations in horse stalls. Journal of Equine Veterinary Science 20(3):197200. Reece, F.N., B.J. Bates, and B.D. Lott. 1979. Ammonia control in broiler houses. Poultry Science 58:754-755.

37

Reece, F.N., B.D. Lott, and W. Deaton. 1980. Ammonia in the atmosphere during brooding affects performance of broiler chickens. Poultry Science 59(3):486-488. Roelofs, J.G.M. and A.L.F.M. Houdijk. 1991. Ecological effects of ammonia. In Cost 681 Expert Odours Group Workshop: Odour and Ammonia Emissions from Livestock Farming, eds, V.C. Nielsen et al. 10-16. Silsoe, UK, 26-29 March 1990. Elsevier Applied Science, London/New York. Rom, H.B. 1993. Ammonia emission from livestock buildings in Denmark. In Fourth International Symposium of Livestock Environment, eds, E. Collins and C. Boon. 1161-1168. University of Warwick, Coventry, England, 6-9 July. ASAE, Michigan. Schmidt-Van Riel, C.J.M. 1991. Experiences with automatic ammonia analyses. In Cost 681 Expert Odours Group Workshop: Odour and Ammonia Emissions from Livestock Farming, eds, V.C. Nielsen et al. 67-70. Silsoe, UK, 26-29 March 1990. Elsevier Applied Science, London. Scholtens, R. 1990. Ammonia emission measurements in animal housing with forced ventilation. In Ammoniak in der Umwelt. Kreislufe, Minderung. Proceedings of a conference, eds, H. Dhler and H. van den Weghe. 9. Braunschweig, Germany, 10-12 October. Scholtens, R. 1993. Gas detector tubes for ammonia Drger and Kitagawa (Gasdetectiebuisjes voor ammoniak: Drger en Kitagawa). Instituut voor Mechanisatie, Arbeid en Gebouwen, Dienst Landbouwkundig Onderzoek. Nota V 93-54. 1-27 p. Secrest, C.D. 2000. Field measurement of air pollutants near swine confined animal feeding operations using UV DOAS and FTIR. In International Optical Engineering Society 2000. Secrest, C.D. 2001. Personal communication. Seltzer, W., S.G. Moum, and T.M. Goldhaft. 1969. A method of treatment of animal waste to control ammonia and other odours. Poultry Science 48:1912-1918. Skewes, P.A. and J.D. Harmon. 1995. Ammonia quick test and ammonia dosimeter tubes for determining ammonia levels in broiler facilities. Journal of Applied Poultry Research 4:148-153. Slanina, S. 1994. Forest dieback and ammonia - a typical Dutch problem. Chemistry International 16:2-3. Snell, H.G.J. and H. Van den Weghe. 1999. Reduction of ammonia emissions and optimization of the stable air temperature by using earth-tube heat exchangers. ASAE Meeting Paper No. 994133. St. Joseph, Mich.: ASAE. Stern, A.C., R.W. Boubel, C.B. Turner, and D.L. Fox. 1984. Fundamentals of air pollution. Second Edition ed. San Diego: Academic Press Inc. 530 p. Stowell, R.R. and S. Foster. 2000. Ammonia emisions from a High-Rise swine finishing facility. ASAE Meeting Paper No. 004080. St. Joseph, Mich.: ASAE. Stowell, R.R., H. Keener, D. Elwell, T. Menke, and S. Foster. 2000. High-RiseTM Hog Facility. In Swine Housing. 273-282. Des Moines, Iowa, October 9-11. St. Joseph, Mich.: ASAE. Svensson, L. and M. Ferm. 1993. Mass transfer coefficient and equilibrium concentration as key factors in a new approach to estimate ammonia emission from livestock manure. Journal of Agricultural Engineering Research 56:1-11. Svensson, L., K.-H. Jeppsson, and G. Gustafsson. 1997. Evaluation of different methods of measuring NH3 emission in naturally ventilated animal houses with deep litter. In International Symposium on Ammonia and Odour Control from Animal Production Facilities, eds, J.A.M. Voermans and G.J. Monteny. Vol. I. 209-217. Vinkeloord, The Netherlands, 6-10 Oct. Rosmalen, The Netherlands: NVTL.

38

TEI. 1995. Model 17 Chemiluminescence NH3 Analyzer Instruction Manual. Thermo Environmental Instruments Inc., Franklin, Massachusetts. P/N 7763. April. USEPA. 1997. EPA Traceability protocol for assay and certification of gaseous calibration standards. National Exposure Research Laboratory, Human Exposure and Atmospheric Science Division, U.S. Environmental Protection Agency, Research Triangle Park, NC. EPA-600/R-97/121. September. USEPA. 1998. EPA Guidance for Quality Assurance Project Plans. Office of Research and Development, United States Environmental Protection Agency, Washington, D.C. EPA QA/G-5. February. Valentine, H. 1964. A study of the effect of different ventilation rates on the ammonia concentrations in the atmosphere of broiler houses. British Poultry Science 5:149-159. Valli, L., S. Piccinini, and G. Bonazzi. 1991. Ammonia emission from two poultry manure drying systems. In Cost 681 Expert Odours Group Workshop: Odour and Ammonia Emissions from Livestock Farming, eds, V.C. Nielsen et al. 50-58. Silsoe, UK, 26-29 March. Elsevier Applied Science, London/New York. van Breemen, N., P.A. Burrough, E.J. Velthorst, H.F. van Dobben, Toke de Wit, T.B. Ridder, and H.F.R. Reijnders. 1982. Soil acidification from atmospheric ammonium sulphate in forest canopy throughfall. Nature 299(20):548-550. Van der Hoek, K.W. 1998. Estimating ammonia emission factors in Europe: summary of the work of the UNECE ammonia expert panel. Atmospheric Environment 32(3):315-316. van Ouwerkerk, E.N.J. 1993. Methods of measuring NH3 emission from animal housing (Meetmethoden Ammoniak Emissie uit Stallen). DOL, IMAG, Wageningen. No. 16. 178 p. van 't Klooster, C.E. and B.P. Heitlager. 1992. Measurement systems for emissions of ammonia and other gasses at the Research Institute for Pig Husbandry. Research Institute for Pig Husbandry, The Netherlands. P3.92. 22 p. Verdoes, N. and N.W.M. Ogink. 1997. Odour emission from pig houses with low ammonia emission. In International Symposium on Ammonia and Odour Control from Animal Production Facilities, eds, J.A.M. Voermans and G.J. Monteny. Vol. I. 317-325. Vinkeloord, The Netherlands, 6-10 Oct. Rosmalen, The Netherlands: NVTL. Verstegen, W.A., W. Van der Hel, A.A. Jongebreur, and G. Enneman. 1976. The influence of ammonia and humidity on activity and energy balance data in groups of pigs. Z.Tierphysilo., TierernShrg.u.Futtermittelkde 37:255-263. Wang, Z., O. Van Cleemput, P. Demeyer, and L. Baert. 1991. Effect of urease inhibitors on urea hydrolysis and ammonia volatilization. Biology and Fertility of Soils(11):43-47. Wheeler, E.F., D.E. Buffington, W. Weaver, R.M. Hulet, J. Shufran, and R. Weiss. 1998. Winter ammonia control in Pennsylvania Broiler Barns, a preliminary analysis of one study farm. ASAE meeting 984039. St. Joseph, Mich.: ASAE. Wheeler, E.F., J.L. Smith, and R.M. Hulet. 2000a. Ammonia volatilization from litter during nine broiler flocks. In Air Pollution from Agricultural Operations, Proceedings of the Second International Conference. 25-32. Des Moines, Iowa, October 9-11. ASAE. Wheeler, E.F., R.W.J. Weiss, and E. Weidenboerner. 1999. Evaluation of instrumentation for the measurement of aerial ammonia in poultry houses. ASAE Meeting Paper No. 993188. St. Joseph, Mich.: ASAE. Wheeler, E.F., R.W.J. Weiss, and E. Weidenboerner. 2000b. Evaluation of instrumentation for measuring aerial ammonia in poultry houses. Journal of Applied Poultry Research 9(4):443-452.

39

White, R. 1990. Chromatography/Fourier transform infrared spectroscopy and its applications. vol. 10. New York: M. Dekker. 328 p. Xin, H., I.L. Berry, and G.T. Tabler. 1996. Minimum ventilation requirement and associated energy cost for aerial ammonia control in broiler houses. Transactions of the ASAE 39(2):645-648.

40

You might also like