Tải bản đầy đủ (.pdf) (32 trang)

RELATIONAL MANAGEMENT and DISPLAY of SITE ENVIRONMENTAL DATA - PART 3 potx

Bạn đang xem bản rút gọn của tài liệu. Xem và tải ngay bản đầy đủ của tài liệu tại đây (421.16 KB, 32 trang )

PART THREE - GATHERING
ENVIRONMENTAL DATA
© 2002 by CRC Press LLC
CHAPTER 10
SITE INVESTIGATION AND REMEDIATION
The site investigation and remediation process is usually the reason for site environmental data
management. The results of the data management process can provide vital input in the decision-
making process. This chapter provides an overview of the regulations that drive the site
investigation and remediation process, some information on how the process works under the
major environmental regulations, and how data management and display is involved in the
different parts of the process. Related processes are environmental assessments and environmental
impact statements, which can also be aided by an EDMS.
OVERVIEW OF ENVIRONMENTAL REGULATIONS
The environmental industry is driven by government regulations. These regulations have been
enacted at the national, state, or local level. Nearly all environmental investigation and remediation
activity is performed to satisfy regulatory requirements. A good overview of environmental
regulations can be found in Mackenthun (1998). The following are some of the most significant
environmental regulations:
National Environmental Policy Act of 1969 (NEPA) – Requires federal agencies to consider
potentially significant environmental impacts of major federal actions prior to taking the action.
The NEPA process contains three levels of possible documentation: 1) Categorical Exclusion
(CATEX), where no significant effects are found, 2) Environmental Assessment (EA), which
addresses various aspects of the project including alternatives, potential impacts, and mitigation
measures, and 3) Environmental Impact Statement (EIS), which covers topics similar to an EA, but
in more detail.
Clean Air Act of 1970 (CAA) – Provides for the designation of air quality control regions,
and requires National Ambient Air Quality Standards (NAAQS) for six criteria pollutants
(particulate matter, sulfur dioxide, carbon monoxide, ozone, nitrogen dioxide, and lead). Also
requires National Emission Standards for Hazardous Air Pollutants (NESHAPs) for 189 hazardous
air pollutants. The act requires states to implement NAAQS, and requires that source performance
standards be developed and attained by new sources of air pollution.


Occupational Safety and Health Act of 1970 – Requires private employers to provide a
place of employment safe from recognized hazards. The act is administered by the Occupational
Safety and Health Administration (OSHA).
© 2002 by CRC Press LLC
Endangered Species Act of 1973 (ESA) – Provides for the listing of threatened or
endangered species. Any federal actions must be evaluated for their impact on endangered species,
and the act makes it illegal to harm, pursue, kill, etc. a listed endangered or threatened species.
Safe Drinking Water Act of 1974 (SDWA) – Protects groundwater aquifers and provides
standards to ensure safe drinking water at the tap. It makes drinking water standards applicable to
all public water systems with at least 15 service connections serving at least 25 individuals.
Requires primary drinking water standards that specify maximum contamination at the tap, and
prohibits certain activities that may adversely affect water quality.
Resource Conservation and Recovery Act of 1976 (RCRA) – Regulates hazardous wastes
from their generation through disposal, and protects groundwater from land disposal of hazardous
waste. It requires criteria for identifying and listing of hazardous waste, and covers transportation
and handling of hazardous materials in operating facilities. The act also covers construction,
management of, and releases from underground storage tanks (USTs). In 1999, 20,000 hazardous
waste generators regulated by RCRA produced over 40 million tons of hazardous waste (EPA,
2001b). RCRA was amended in 1984 with the Hazardous and Solid Waste Amendments (HSWA)
that required phasing out land disposal of hazardous waste.
Toxic Substances Control Act of 1976 (TSCA) – Requires testing of any substance that may
present an unreasonable risk of injury to health or the environment, and gives the EPA authority to
regulate these substances. Covers the more than 60,000 substances manufactured or processed, but
excludes nuclear materials, firearms and ammunition, pesticides, tobacco, food additives, drugs,
and cosmetics.
Clean Water Act of 1977 (CWA) – Based on the Federal Water Pollution Control Act of
1972 and several other acts. Amended significantly in 1987. This act, which seeks to eliminate the
discharge of pollutants into navigable waterways, has provisions for managing water quality and
permitting of treatment technology. Development of water quality standards is left to the states,
which must set standards at least as stringent as federal water quality standards.

Comprehensive Environmental Response, Compensation, and Liability Act of 1980
(CERCLA, Superfund) – Enacted to clean up abandoned and inactive hazardous waste sites.
Creates a tax on the manufacture of certain chemicals to create a trust fund called the Superfund.
Sites to be cleaned up are prioritized as a National Priority List (NPL) by the EPA. Procedures and
cleanup criteria are specified by a National Contingency Plan. The NPL originally contained 408
sites, and now contains over 1300. Another 30,000 sites are being evaluated for addition to the list.
Emergency Planning and Community Right-to-Know Act of 1986 (EPCRA) – Enacted
after the Union Carbide plant disaster in Bhopal, India in 1984, in which release of methyl
isocyanate from a chemical plant killed 2,000 and impacted the health of 170,000 survivors, this
law requires industrial facilities to disclose information about chemicals stored onsite.
Pollution Prevention Act of 1990 (PPA) – Requires collection of information on source
reduction, recycling, and treatment of listed hazardous chemicals. Resulted in a Toxic Release
Inventory for facilities including amounts disposed of onsite and sent offsite, recycled, and used for
energy recovery.
These regulations have contributed significantly to improvement of our environment. They
have also resulted in a huge amount of paperwork and other expenses for many organizations, and
explain why environmental coordinators stay very busy.
Bad regulations are more likely to be supplemented than repealed.
Rich (1996)
© 2002 by CRC Press LLC
THE INVESTIGATION AND REMEDIATION PROCESS
The details of the site investigation and remediation process vary depending on the regulation
under which the work is being done. Superfund was designed to remedy mistakes in hazardous
waste management made in the past at sites that have been abandoned or where a sole responsible
party cannot be determined. RCRA deals with sites that have viable operators and ongoing
operations. The majority of sites fall into one of these two categories. The rest operate under a
range of regulations through various different regulatory bodies, many of which are agencies in the
various states.
CERCLA
CERCLA (Superfund) gives the EPA the authority to respond to releases or threatened

releases of hazardous substances that may endanger human health and the environment. The three
major areas of enforcement at Superfund sites are: achieving site investigations and cleanups led
by the potentially responsible party (PRP) or parties (PRP lead cleanups, meaning the lead party
on the project is the PRP); overseeing PRP investigation and cleanup activities; and recovering
from PRPs the costs spent by EPA at Superfund cleanups (Fund lead cleanups).
The National Contingency Plan of CERCLA describes the procedures for identification,
evaluation, and remediation of past hazardous waste disposal sites. These procedures are
preliminary assessment and site inspection; Hazard Ranking System (HRS) scoring and National
Priority List (NPL) site listing; remedial investigation and feasibility studies; record of decision;
remedial design and remedial action; construction completion; operation and maintenance; and
NPL site deletion. Site environmental data can be generated at various steps in the process.
Additional information on Superfund enforcement can be found in EPA (2001a).
Preliminary assessment and site inspection – The process starts with investigations of site
conditions. A preliminary assessment (PA) is a limited scope investigation performed at each site.
Its purpose is to gather readily available information about the site and surrounding area to
determine the threat posed by the site. The site inspection (SI) provides the data needed for the
hazard ranking system, and identifies sites that enter the NPL site listing process (see below). SIs
typically involve environmental and waste sampling that can be managed using the EDMS.
HRS scoring and NPL site listing – The hazard ranking system (HRS) is a numerically
based screening system that uses information from initial, limited investigations to assess the
relative potential of sites to pose a threat to human health or the environment. The HRS assigns a
numerical score to factors that relate to risk based on conditions at the site. The four risk pathways
scored by HRS are groundwater migration; surface water migration; soil exposure; and air
migration.
HRS is the principal mechanism EPA uses to place uncontrolled waste sites on the National
Priorities List (NPL). Identification of a site for the NPL helps the EPA determine which sites
warrant further investigation, make funding decisions, notify the public, and serve notice to PRPs
that EPA may begin remedial action.
Remedial investigation and feasibility studies – Once a site is on the NPL, a remedial
investigation/feasibility study (RI/FS) is conducted at the site. The remedial investigation involves

collection of data to characterize site conditions, determine the nature of the waste, assess the risk
to human health and the environment, and conduct treatability testing to evaluate the potential
performance and cost of the treatment technologies that are being considered. The feasibility study
is then used for the development, screening, and detailed evaluation of alternative remedial actions.
The RI/FS has five phases: scoping; site characterization; development and screening of
alternatives; treatability investigations; and detailed analyses. The EDMS can make a significant
contribution to the site characterization component of the RI/FS, which often involves a significant
amount of sampling of soil water and air at the site The EDMS serves as a repository of the data
© 2002 by CRC Press LLC
as well as a tool for data selection and analysis to support the decision-making process. Part of the
site characterization process is to develop a baseline risk assessment to identify the existing or
potential risks that may be posed to human health and environment at the site. The EDMS can be
very useful in this process by helping screen the data for exceedences that may represent risk
factors.
Record of decision – Once the RI/FS has been completed, a record of decision (ROD) is
issued that explains which of the cleanup alternatives will be used to clean up the site. This public
document can be significant for data management activities because it often sets target levels for
contaminants that will be used in the EDMS for filtering, comparison, and so on.
Remedial design and remedial action – In the remedial design (RD), the technical
specifications for cleanup remedies and technologies are designed. The remedial action (RA)
follows the remedial design and involves the construction or implementation phase of the site
cleanup. The RD/RA is based on specifications described in the ROD. The EDMS can assist
greatly with tracking the progress of the RA and determining when ROD limits have been met.
Construction completion – A construction completion list (CCL) helps identify successful
completion of cleanup activities. Sites qualify for construction completion when any physical
construction is complete (whether or not cleanup levels have been met), EPA has determined that
construction is not required, or the site qualifies for deletion from the NPL.
Operation and maintenance – Operation and maintenance (O&M) activities protect the
integrity of the selected remedy for a site, and are initiated by the state after the site has achieved
the actions and goals outlined in the ROD. The site is then determined to be operational and

functional (O&F) based on state and federal agreement when the remedy for a site is functioning
properly and performing as designed, or has been in place for one year. O&M monitoring involves
inspection; sampling and analysis; routine maintenance; and reporting. The EDMS is used heavily
in this stage of the process.
NPL site deletion – In this final step, sites are removed from the NPL once they are judged to
no longer be a significant threat to human health and the environment. To date, not many sites have
been delisted.
RCRA
The EPA’s Office of Solid Waste (OSW) is responsible for ensuring that currently generated
solid waste is managed properly, and that currently operating facilities address any contaminant
releases from their operations. In some cases, accidents or other activities at RCRA facilities have
released hazardous materials into the environment, and the RCRA Corrective Action Program
covers the investigation and cleanup of these facilities. Additional information on RCRA
enforcement can be found in EPA (2001b).
As a condition of receiving a RCRA operating permit, active facilities are required to clean up
contaminants that are being released or have been released in the past. EPA, in cooperation with
the states, verifies compliance through compliance monitoring, educational activities, voluntary
incentive programs, and a strong enforcement program. The EDMS is heavily involved in
compliance monitoring and to some degree in enforcement actions.
Compliance monitoring – EPA and the states determine a waste handler’s compliance with
RCRA requirements using inspections, record reviews, sampling, and other activities. The EDMS
can generate reports comparing sampling results to regulatory limits to save time in the compliance
monitoring process.
Enforcement actions – The compliance monitoring process can turn up violations, and
enforcement actions are taken to bring the waste handler into compliance and deter further
violations. These actions can include administrative actions, civil judicial actions, and criminal
actions. In addition, citizens can file suit to bring enforcement actions against violators or potential
violators.
© 2002 by CRC Press LLC
One important distinction from a data management perspective between CERCLA and RCRA

projects is that CERCLA projects deal with past processes, while RCRA projects deal with both
past and present processes. This means that the EDMS for both projects needs to store information
on soil, groundwater, etc., while the RCRA EDMS also might store information on ongoing
processes such as effluent concentrations and volumes, and even production and other operational
information.
Other regulatory oversight
While many sites are investigated and remediated under CERCLA or RCRA, other regulatory
oversight is also possible. The EPA has certified some states to oversee cleanup within their
boundaries. In some cases, other government agencies, including the armed forces, oversee their
own cleanup efforts. In general, the technical activities performed are pretty much the same
regardless of the type of oversight, and the functional requirements for the EDMS are also the same
The main exception is that some of these agencies require the use of specific reporting tools as
described in Chapter 5.
ENVIRONMENTAL ASSESSMENTS AND ENVIRONMENTAL
IMPACT STATEMENTS
The National Environmental Policy Act of 1969 (NEPA), along with various supplemental
laws and legal decisions, requires federal agencies to consider the environmental impacts and
possible alternatives of any federal actions that significantly affect the environment (Mackenthun,
1998, p. 15; Yost, 1997, p. 1-11). This usually starts with an environmental assessment (EA). The
EA can result in a determination that an environmental impact statement (EIS) is required, or in a
finding of no significant impact (FONSI). The EIS is a document that is prepared to assist with
decision making based on the environmental consequences and reasonable alternatives of the
action. The format of an EIS is recommended in 40 CFR 1502.10, and is normally limited to 150
pages. Often there is considerable public involvement in this process.
One important use of environmental assessments is in real estate transactions. The seller and
especially the buyer want to be aware of any environmental liabilities related to the property being
transferred. These assessments are broken into phases. The data management requirements of EAs
and EISs vary considerably, depending on the nature of the project and the amount and type of data
available.
Phase 1 Environmental Assessment – This process involves evaluation of existing data

about a site, along with a visual inspection, followed by a written report, similar to a preliminary
assessment and site inspection under CERCLA, and can satisfy some CERCLA requirements such
as the innocent landowner defense. The Phase 1 assessment process is well defined, and guidelines
such as Practice E-1527-00 from the American Society for Testing and Materials (ASTM 2001a,
2001b), are used for the assessment and reporting process. There are four parts to this process:
gathering information about past and present activities and uses at the site and adjoining properties;
reviewing environmental files maintained by the site owner and regulatory agencies; inspection of
the site by an environmental professional; and preparation of a report identifying existing and
potential sources of contamination on the property. The work involves document searches and
review of air photos and site maps. Often the source materials are in hard copy not amenable to
data management. Public and private databases are available to search ownership, toxic substance
release, and other information, but this data is usually managed by its providers and not by the
person performing the search. Phase 1 assessments for a small property are generally not long or
complicated, and can cost as little as $1,000.
© 2002 by CRC Press LLC
Phase 2 Investigation – If a Phase 1 assessment determines that the presence of contam-
ination is likely, the next step is a Phase 2 assessment. The primary differences are that Phase 1
relies on existing data, while in Phase 2 new data is gathered, usually in an intrusive manner, and
the Phase 2 process is less well defined. This can involve sampling soil, sediment, and sludge and
installation of wells for sampling groundwater. This is similar to remedial investigation and
feasibility studies under CERCLA. If the assessment progresses to the point where samples are
being taken and analyzed, then the in-house data management system can be of value.
Phase 3 Site Remediation and Decommissioning – The final step of the assessment process,
if necessary, is to perform the cleanup and assess the results. Motivation for the remediation might
include the need to improve conditions prior to a property transfer, to prevent contamination from
migrating off the property, to improve the value of the property, or to avoid future liability.
Monitoring the cleanup process, which can involve ongoing sampling and analysis, will usually
involve the EDMS.
© 2002 by CRC Press LLC
CHAPTER 11

GATHERING SAMPLES
AND DATA IN THE FIELD
Environmental monitoring at industrial and other facilities can involve one or more different
media. The most common are soil, sediment, groundwater, surface water, and air. Other media of
concern in specific situations might include dust, paint, waste, sludge, plants and animals, and
blood and tissue. Each medium has its own data requirements and special problems. Generating
site environmental data starts with preparing sampling plans and gathering the samples and related
data in the field. There are a number of aspects of this process that can have a significant impact on
the resulting data quality. Because the sampling process is specific to the medium being sampled,
this chapter is organized by medium. Only the major media are discussed in detail.
GENERAL SAMPLING ISSUES
The process of gathering data in the field, sending samples to the laboratory, analyzing the
data, and reporting the results is complicated and error-prone. The people doing the work are often
overworked and underpaid (who isn’t), and the requirements to do a good job are stringent.
Problems that can lead to questionable or unusable data can occur at any step of the way. The
exercise (and in some cases, requirement) of preparing sampling plans can help minimize field data
problems. Field sampling activities must be fully documented in conformance with project quality
guidelines. Those guidelines should be carefully thought out and followed methodically. A few
general issues are covered here. The purpose of this section is not to teach field personnel to
perform the sampling, but to help data management staff understand where the samples and data
come from in order to use it properly. In all cases, regulations and project plans should be followed
in preference to statements made here. Additional information on these issues can be found in
ASTM (1997), DOE/HWP (1990a), and Lapham, Wilde, and Koterba (1985).
Taking representative samples
Joseph (1998) points out that the basic premise of sampling is that the sample must represent
the whole population, and quotes the law of statistical regularity as stating that “a set of subjects
taken at random from a large group tends to reproduce the characteristics of that large group.” But
the sample is only valid if the errors introduced in the sampling process do not invalidate the
results for the purpose intended for the samples. Analysis of the samples should result in no bias
and minimum random errors.

© 2002 by CRC Press LLC
Types of Sampling Patterns
Simple Random Sampling
Judgment Sampling
Grid (Systematic) Sampling
Stratified Sampling
Random Grid Sampling
Two-Stage Sampling
Known
Plume
Primary
Stage
Secondary
Stage
BA
Figure 51 - Types of sampling patterns
The size of the sample set is directly related to the precision of the result. More samples cost
more money, but give a more reliable result. If you start with the precision required, then the
number of samples required can be calculated:
© 2002 by CRC Press LLC
22
2
)4/( sBN
Ns
n
+
=
where n is the number of samples, N is the size of the population, s is the standard deviation of the
sample, and B is the desired precision, such as 95% confidence. According to Joseph (1998), the
standard deviation can be estimated by taking the largest value of the data minus the smallest and

dividing by four.
There are several strategies for laying out a sampling program. Figure 51, modified after
Adolfo and Rosecrance (1993), shows six possibilities. Sampling strategies are also discussed in
Sara (1994, p. 10-49). In simple random sampling, the chance of selecting any particular location is
the same. With judgment sampling, sampling points are selected based on previous knowledge of
the system to be sampled. Grid sampling provides uniform coverage of the area to be studied.
Stratified sampling has the sample locations based on existence of discrete areas of interest, such as
aquifers and confining layers, or disposal ponds and the areas between them. Random grid
sampling combines uniform coverage of the study area with a degree of random selection of each
location, which can be useful when access to some locations is difficult. With two-stage sampling,
secondary sampling locations are based on results of primary stage samples. In the example shown,
primary sample A had elevated values, so additional samples were taken nearby, while primary
sample B was clean, so no follow-up samples were taken.
Care should be taken so that the sample locations are as representative as possible of the
conditions being investigated. For example, well and sample locations near roadways may be
influenced by salting and weed spraying activities. Also, cross-contamination from dirty samples
must be avoided by using procedures like sampling first from areas expected to have the least
contamination, then progressing to areas expected to have more.
Logbooks and record forms
Field activities must be fully documented using site and field logbooks. The site logbook
stores information on all field investigative activities, and is the master record of those activities.
The field logbook covers the same activities, but in more detail. The laboratory also should keep a
logbook for tracking the samples after they receive them.
The field logbook should be kept up-to-date at all times. It should include information such as
well identification; date and time of sampling; depth; fluid levels; yield; purge volume, pumping
rate, and time; collection methods; evacuation procedures; sampling sequence; container types and
sample identification numbers; preservation; requested parameters; field analysis data; sample
distribution and transportation plans; name of collector; and sampling conditions.
Several field record forms are used as part of the sampling process. These include Sample
Identification and Chain of Custody forms. Also important are sample seals to preserve the

integrity of the sample between sampling and when it is opened in the laboratory. These are legal
documents, and should be created and handled with great care.
Sample Identification forms are usually a label or tag so that they stay with the sample. Labels
must be waterproof and completed in permanent ink. These forms should contain such information
as site name; unique field identification of sample, such as station number; date and time of sample
collection; type of sample (matrix) and method of collection; name of person taking the sample;
sample preservation; and type of analyses to be conducted.
Chain of Custody (COC) forms make it possible to trace a sample from the sampling event
through transport and analysis. The COC must contain the following information: project name;
signature of sampler; identification of sampling stations; unique sample numbers; date and time of
collection and of sample possession; grab or composite designation; matrix; number of containers;
parameters requested for analysis; preservatives and shipping temperatures; and signatures of
individuals involved in sample transfer.
© 2002 by CRC Press LLC
COC forms should be enclosed in a plastic cover and placed in the shipping container with the
samples. When the samples are given to the shipping company, the shipping receipt number should
be recorded on the COC and in the site logbook. All transfers should be documented with the
signature, date, and time on the form. A sample must remain under custody at all times. A sample
is under custody if it is in the sampler’s possession; it is in the sampler’s view after being in
possession; it is in the possession of a traceable carrier; it is in the possession of another
responsible party such as a laboratory; or it is in a designated secure area.
SOIL
Taking soil samples must take into consideration that soil is a very complex physical material.
The solid component of soil is a mix of inorganic and organic materials. In place in the ground,
soil can contain one or more liquid phases and a gas phase, and these can be absorbed or adsorbed
in various ways. The sampling, transportation, and analysis processes must be managed carefully
so that analytical results accurately represent the true composition of the sample.
Soil sampling issues
Before a soil sample can be taken, the material to be sampled must be exposed. For surface or
shallow subsurface soil samples this is generally not an issue, but for subsurface samples this

usually requires digging. This can be done using either drilling or drive methods. For
unconsolidated formations, drilling can be done using an auger (hollow-stem, solid flight, or
bucket), drilling (rotary, sonic, directional), or jetting methods. For consolidated formations, rotary
drilling (rotary bit, downhole hammer, or diamond drill) or cable tools can be used. Drive methods
include cone penetrometers or direct push samplers.
Sometimes it is useful to do a borehole geophysical survey after the hole is drilled. Examples
of typical measurements include spontaneous potential, resistivity, gamma and neutron surveys,
acoustic velocity, caliper, temperature, fluid flow, and electromagnetic induction.
Soil samples are gathered with a variety of tools, including spoons, scoops, shovels, tubes, and
cores. The samples are then sealed and sent to the laboratory. Duplicates should be taken as
required by the QAPP (quality assurance project plan).
Sometimes soil samples are taken as a boring is made, and then the boring is converted to a
monitoring well for groundwater, so both soil and water samples may come from the same hole in
the ground.
Typical requirements for soil samples are as follows. The collection points should be surveyed
relative to a permanent reference point, located on a site map, and referenced in the field logbook.
A clean, decontaminated auger, spoon, or trowel should be used for each sample collected. Surface
or air contact should be minimized by placing the sample in an airtight container immediately after
collection. The sampling information should be recorded in the field logbook and any other
appropriate forms.
For subsurface samples, the process for verifying depth of sampling, the depth control
tolerance, and the devices used to capture the samples should be as specified in the work plan. Care
must be taken to prevent cross-contamination or misidentification of samples.
Sometimes the gas content of soil is of concern, and special sampling techniques must be used.
These include static soil gas sampling, soil gas probes, and air sampling devices.
Velilind’s laws of Experimentation: 1. If reproducibility may be a problem, conduct the test
only once. 2. If a straight line fit is required, obtain only two data points.
McMullen (1996)
© 2002 by CRC Press LLC
Special consideration should be given for soil and sediment samples to be analyzed for volatile

organics (VOAs). The samples should be taken with the least disturbance possible, such as using
California tubes. Use Teflon or stainless steel equipment. If preservatives are required, they should
be added to the bottle before sampling. Samples for VOA analysis should not be split. Air bubbles
cannot be present in the sample. The sample should never be frozen.
Soil data issues
Soil data is usually gathered in discrete samples, either as surface samples or as part of a soil
boring or well installation process. Then the sample is sent to the laboratory for analysis, which can
be chemical, physical, or both. Each sample has a specific concentration of each constituent of
concern. Sometimes it is useful to know not only the concentration of a toxin, but also its mobility
in groundwater. Useful information can be provided by a leach test such as TCLP (toxicity
characterization leaching procedure), in which a liquid is passed through the soil sample and the
concentration in the leachate is measured. This process is described in more detail in Chapter 12.
Key data fields for soil samples include the site and station name; the sample date and depth;
COC and other field sample identification numbers; how the sample was taken and who took it;
transportation information; and any sample QC information such as duplicate and other
designations. For surface soil samples, the map coordinates are usually important. For subsurface
soil samples, the map coordinates of the well or boring, along with the depth, often as a range (top
and bottom), should be recorded. Often a description of the soil or rock material is recorded as the
sample is taken, along with stratigraphic or other geologic information, and this should be stored in
the EDMS as well.
SEDIMENT
Procedures for taking sediment samples are similar to those for soil samples. Samples should
be collected from areas of least to greatest contamination, and from upstream to downstream.
Sediment plumes and density currents should be avoided during sample collection.
GROUNDWATER
Groundwater is an important resource, and much environmental work involves protecting and
remediating groundwater. A good overview of groundwater and its protection can be found in
Conservation Technology Resource Center (2001). Groundwater accounts for more than 95% of all
fresh water available for use, and about 40% of river flow depends on groundwater. About 50% of
Americans (and 95% of rural residents) obtain all or part of their drinking water from groundwater.

Groundwater samples are usually taken at a location such as a monitoring well, for an
extended period of time such as quarterly, for many years. Additional information on groundwater
sampling can be found in NWWA (1986) and Triplett (1997).
Groundwater or Ground Water?
Is “groundwater” one word or two? When used by itself, groundwater as one word looks fine,
and many people write it this way. The problem comes in when it is written along with
surface water, which is always two words, and ground water as two words looks better. Some
individuals and organizations prefer it one way, some the other, so apparently neither is right
or wrong. For any one writer, just as with “data is” vs. “data are,” the most important thing is
to pick one and be consistent.
© 2002 by CRC Press LLC
Figure 52 - Submersible sampling pump (Courtesy of Geotech Environmental Equipment)
Groundwater sampling issues
The first step in groundwater sampling is to select the location and drill the hole. Drilling
methods are similar to those described above in the section on soil sampling, and soil samples can
be taken when a groundwater well is drilled. Then the wellbore equipment such as tubing, screens,
and annular material is placed in the hole to make the well. The tubing closes off part of the hole
and the screens open the other part so water can enter the wellbore. Screening must be at the
correct depth so the right interval is being sampled.
Prior to the first sampling event, the well is developed. For each subsequent event it is purged
and then sampled. The following discussion is intended to generally cover the issues of
groundwater sampling. The details of the sampling process should be covered in the project work
plan. Appropriate physical measurements of the groundwater are taken in the field. The sample is
placed in a bottle, preserved as appropriate, chilled and placed in a cooler, and sent to the
laboratory.
Well development begins sometime after the well is installed. A 24-hour delay is typical.
Water is removed from the well by pumping or bailing, and development usually continues until
the water produced is clear and free of suspended solids and is representative of the geologic
formation being sampled. Development should be documented on the Well Development Log
Form and in the site and field logbooks. Upgradient and background wells should be developed

before downgradient wells to reduce the risk of cross-contamination.
Measurement of water levels should be done according to the sampling plan, which may
specify measurement prior to purging or prior to sampling. Groundwater level should be measured
to a specific accuracy (such as 0.05 ft) and with a specific precision (such as 0.01 ft).
Measurements should be made relative to a known, surveyed datum. Measurements are taken with
a steel tape or an electronic device such as manometer or acoustical sounder. Some wells have a
pressure transducer installed so water levels can be obtained more easily.
© 2002 by CRC Press LLC
Figure 53 - Multi-parameter field meter (Courtesy of Geotech Environmental Equipment)
Some wells contain immiscible fluid layers, called non-aqueous phase liquids (NAPLs). There
can be up to three layers, which include the water and lighter and heavier fluids.
The lighter fluids, called light non-aqueous phase liquids (LNAPLs) or floaters, accumulate
above the water. The heavier fluids, called dense non-aqueous phase liquids (DNAPLs) or sinkers,
accumulate below the water. For example, LNAPLs like gasoline float on water, while DNAPLs
such as chlorinated hydrocarbons (TCE, TCA, and PCE) sink. NAPLs can have their own flow
regime in the subsurface separate from the groundwater. The amount of these fluids should be
measured separately, and the fluids collected, prior to purging. For information on measurement of
DNAPL, see Sara (1994, p. 10-75).
Purging is done to remove stagnant water in the casing and filter pack so that the water
sampled is “fresh” formation water. A certain number of water column volumes (such as three) are
purged, and temperature, pH, and conductivity must be monitored during purging to ensure that
these parameters have stabilized prior to sampling. Upgradient and background wells should be
purged and sampled before downgradient wells to reduce the risk of cross-contamination.
Information concerning well purging should be documented in the Field Sampling Log.
Sampling should be done within a specific time period (such as three hours) of purging, if
recharge is sufficient, otherwise as soon as recharge allows. The construction materials of the
sampling equipment should be compatible with known and suspected contaminants. Groundwater
sampling is done using various types of pumps including bladder, gear, submersible rotor,
centrifugal, suction, or inertial lift; or with a bailed rope. Pumping is usually preferred over bailing
because it is takes less effort and causes less disturbance in the wellbore. An example of a

submersible pump is shown in Figure 52.
Field measurements should be taken at the time of sampling. These measurements, such as
temperature, pH, and specific conductance, should be taken before and after the sample is collected
to check on the stability of the water during sampling. Figure 53 shows a field meter for taking
these measurements. The field data (also known as the field parameters) is entered on the COC,
and should be entered into the EDMS along with the laboratory analysis data. Sometimes the
laboratory will enter this data for the client and deliver it with the electronic data deliverable.
© 2002 by CRC Press LLC
At all stages of the sampling process it is critical that contamination of the samples be
prevented. Contamination can be minimized by properly storing and transporting sampling
equipment, keeping equipment and bottles away from sources of contamination, using clean hands
and gloves to handle equipment and bottles, and carefully cleaning the purging and sampling
equipment after use.
If sampling is for VOAs (volatile organic analysis) then equipment or processes that can
agitate and potentially volatilize samples should be avoided. Sampling methods such as bottom-
filling bailers of stainless steel or Teflon and/or Teflon bladder pumps should be used.
Powell and Puls (1997) have expressed a concern that traditional groundwater sampling
techniques, which are largely based on methods developed for water supply investigations, may not
correctly represent the true values or extent of a plume. For example, the turbidity of a sample is
often related to the concentration of constituents measured in the sample, and sometimes this may
be due to sampling methods that cause turbulence during sampling, resulting in high concentrations
not representative of in-situ conditions. Filtering the sample can help with this, but a better
approach may be to use sampling techniques that cause less disturbance of materials in the
wellbore. Small diameter wells, short screened intervals, careful device insertion (or the use of
permanently installed devices), and low pump rates (also known as low-flow samples) are
examples of techniques that may lead to more representative samples.
Preservation and handling of the samples is critical for obtaining reliable analytical results.
Groundwater samples are usually treated with a preservative such as nitric, sulfuric, or
hydrochloric acid or sodium hydroxide (depending on the parameter) to stabilize the analytes, and
then cooled (typically to 4°C) and shipped to the laboratory. The shipping method is important

because the analyses must be performed within a certain period (holding time) after sampling. The
preservation and shipping process varies for different groups of analytes. See Appendix D for more
information about this.
Groundwater data issues
The sample is taken and often some parameters are measured in the field such as temperature,
pH, and turbidity. Then the sample is sent to the laboratory for analysis. When the field and
laboratory data are sent to the data administrator, the software should help the data administrator
tie the field data to the laboratory data for each sampling event.
Key data fields for groundwater data include the site and station name, the sample date and
perhaps time, COC and other field sample identification numbers, how the sample was taken and
who took it, transportation information, and any sample QC information such as duplicate and
other designations. All of this data should be entered into the EDMS.
SURFACE WATER
Surface water samples have no purging requirements, but are otherwise sampled and
transferred the same as groundwater samples. Surface water samples may be easier to acquire, so
they may be taken more often than groundwater samples.
Surface water sampling issues
Surface water samples can be taken either at a specific map location, or at an appropriate
location and depth. The location of the sample should be identified on a site map and described in
the field logbook. Samples should progress from areas of least contamination to worst
contamination and generally from upstream to downstream. The sample container should be
submerged with the mouth facing upstream (to prevent bubbles in the sample), and sample
© 2002 by CRC Press LLC
information should be recorded in the field logbook and any other appropriate forms. The devices
used and the process for verifying depth and depth control tolerance should be as specified in the
project work plan.
Surface water data issues
The data requirements for surface water samples are similar to groundwater samples. For
samples taken in tidal areas, the status of the tide (high or low) should be noted.
DECONTAMINATION OF EQUIPMENT

Equipment must be decontaminated prior to use and re-use. The standard operating procedure
for decontamination should be in the project work plan. The decontamination process is usually
different for different equipment. The following are examples of equipment decontamination
procedures (DOE/HWP 1990a). For nonsampling equipment such as rigs, backhoes, augers, drill
pipe, casing, and screen, decontaminate with high pressure steam, and if necessary scrub with
laboratory-grade detergent and rinse with tap water. For sampling equipment used in inorganic
sampling, scrub with laboratory-grade detergent, rinse with tap water, rinse with ASTM Type II
water, air-dry, and cover with plastic sheeting. For sampling equipment used in inorganic or
organic sampling, scrub with laboratory-grade detergent, rinse with tap water, rinse with ASTM
Type II water, rinse with methanol (followed by a hexane rinse if testing for pesticides, PCBs, or
fuels), air-dry, and wrap with aluminum foil.
SHIPPING OF SAMPLES
Samples should be shipped in insulated carriers with either freezer forms (“blue ice”) or wet
ice. If wet ice is used, it should be placed in leak-proof plastic bags. Shipping containers should be
secured with nylon reinforced strapping tape. Custody seals should be placed on the containers to
verify that samples are not disturbed during transport. Shipping should be via overnight express
within 24 hours of collection so the laboratory can meet holding time requirements.
AIR
The data management requirements for air sampling are somewhat different from those of soil
and water because the sampling process is quite different. While both types of data can (and often
should) be stored in the same database system, different data fields may be used, or the same fields
used differently, for the different types of data. As an example, soil samples will have a depth
below ground, while air samples may have an elevation above ground. Typical air quality
parameters include sulfur dioxide, nitrogen dioxide, carbon monoxide, ozone, and lead. Other
constituents of concern can be measured in specific situations. Sources of air pollution include
transportation, stationary fuel combustion, industrial processes, solid waste disposal, and others.
For an overview of air sampling and analysis, see Patnaik (1997). For details on several air
sampling methods, see ASTM (1997).
Air sampling issues
Concentrations of contaminants in air vary greatly from time to time due to weather

conditions, topography, and changes in source input. Weather conditions that may be important
© 2002 by CRC Press LLC
include wind conditions, temperature, humidity, barometric pressure, and amount of solar
radiation.
The taking and analysis of air samples vary widely depending on project requirements. Air
samples can represent either outdoor air or indoor air, and can be acquired by a variety of means,
both manual and automated. Some samples represent a specific volume of air, while others
represent a large volume of air passed through a filter or similar device. Some air measurements
are taken by capturing the air in a container for analysis, while others are done without taking
samples, using real-time sensors. In all cases, a sampling plan should be established and followed
carefully.
Physical samples can be taken in a Tedlar bag, metal (Summa) canister, or glass bulb. The air
may be concentrated using a cryogenic trap, or compressed using a pump. For organic analysis,
adsorbent tubes may be used. Adsorbent materials typically used include activated charcoal, Tenax
(a porous polymer), or silica gel. Particulate matter such as dust, silica, metal, and carbon particles
are collected using membrane filters. A measured volume of air is pumped through the filter, and
the suspended particles are deposited on the filter. For water-soluble analytes such as acid, alkali,
and some organic vapors, samples can be taken using an impinger, where the air is bubbled
through water, and then the water is analyzed. Toxic gases and vapors can be measured using
colorimetric dosimeters, where a tube containing a paper strip or granulated material is exposed to
the air on one end, and the gas diffuses into the tube and causes the material to change color. The
amount of color change reflects the concentration of the constituent being measured.
Automated samples can be taken using ambient air analyzers. Care should be taken that the air
being analyzed is representative of the area under investigation, and standards should be used to
calibrate the analyzer.
Air data issues
Often air samples are taken at relatively short time intervals, sometimes as short as minutes
apart. This results in a large amount of data to store and manipulate, and an increased focus on
time information rather than just date information. It also increases the importance of data
aggregation and summarization features in the EDMS so that the large volume of data can be

presented in an informative way.
Key data fields for air data include the site and station name, the sample date and time (or, for
a sample composited over time, the start and end dates and times), how the sample was taken and
who took it, transportation information if any, and any sample QC information such as duplicate
and other designations.
OTHER MEDIA
The variety of media that can be analyzed to gather environmental information is almost
unlimited. This section covers just a few of the possible media. There certainly are many others
routinely being analyzed, and more being added over time.
Human tissue and fluids
Exposure to toxic materials can result in the buildup of these materials in the body. Tracking
this exposure involves measuring the concentration of these materials in tissue and body fluids. For
example, hair samples can provide a recent history of exposure, and blood and urine analyses are
widely used to track exposure to metals such as lead, arsenic, and cadmium. Often this type of data
is gathered under patient confidentiality rules, and maintaining this confidentiality must be
considered in implementing and operating the system for managing the data. Lead exposure in
© 2002 by CRC Press LLC
children (and pregnant and nursing mothers) is of special interest since it appears to be correlated
with developmental problems, and monitoring and remediating elevated blood lead is receiving
much attention in some communities. The data management system should be capable of managing
both the blood lead data and the residential environmental data (soil, paint, water, and dust) for the
children. It should also be capable of relating the two even if the blood data is within the patient
confidentiality umbrella and the residential environmental data is not.
Organisms
Because each level of the food chain can concentrate pollutants by one or more orders of
magnitude, the concentration of toxins in biologic material can be a key indicator of those toxins in
the environment. In addition, some organisms themselves can pose a health hazard. Both kinds of
information might need to be stored in a database. Sampling procedures vary depending on the size
of the organisms and whether they are benthic (attached to the bottom), planktonic (move by
floating), or nektonic (move by swimming). For more information, see ASTM (1997).

Residential and workplace media
Increasingly, the environmental quality in homes and offices is becoming a concern. From
toxic materials, to “sick building syndrome” and infectious diseases like anthrax and Legionnaire’s
disease, the quality of the indoor environment is coming into question, and it is logical to track this
information in a database. Other environmental issues relate to exposure to toxic materials such as
lead. Lead can occur in paint, dust, plumbing, soil, and in household accessories, including such
common objects as plastic mini-blinds from the local discount store. Concentration information can
be stored in a database, and the EDMS can be used to correlate human exposure information with
residential or workplace media information to assist with source identification and remediation.
Plant operating and other data
Often information on plant operations is important in management of the environmental issues
at a facility. The relationship between releases and production volume, or chemical composition vs.
physical properties, can be best investigated if some plant operating information is captured and
stored in the EDMS. One issue to keep in mind is that the volume of this information can be great,
and care should be taken to prevent degradation of the performance of the system due to the
potentially large volume of this data. Sometimes it makes sense to store true operating data in one
database and environmental data in another. However, due to the potential overlap of retrieval
requirements, a combined database or duplicated storage for the data with dual uses is sometimes
necessary. Also, the reporting of plant operating data and its relationship to environmental factors
often involves deriving results through calculations. For example, what is measured may be the
volume of effluent and the concentration of some material in the effluent, but perhaps the operating
permit sets limits on the total amount of material. The reporting process needs to multiply the
volume times the concentration to get the amount, and perhaps scale that to different units. Figure
54 shows a screen for defining calculated parameters.
OVERVIEW OF PARAMETERS
The EDMS is used in environmental investigation and remediation projects to track and
understand the amount and location of hazardous materials in the environment. In addition, other
parameters may be tracked to help better understand the amount and distribution of the
contaminants.
© 2002 by CRC Press LLC

Figure 54 - Screen for defining calculated parameters
This section briefly covers some of the environmental and related parameters that are likely to
be managed in an EDMS. Other sources cover this material in much more detail. Useful reference
books include Manahan (2000, 2001), Patnaik (1997), and Weiner (2000). Many Web sites include
reference information on parameters. Examples include EPA (2000a), EXTOXNET (2001),
Spectrum Labs (2001), NCDWQ (2001), SKC (2001), and Cambridge Software (2001). A Web
search will turn up many more. This section covers inorganic parameters, organic parameters, and
various other parameters commonly found in an EDMS.
Inorganic parameters
Inorganic compounds include common metals, heavy metals, nutrients, inorganic nonmetals,
and radiologic parameters.
Common metals include calcium, iron, magnesium, potassium, sodium, and others. These
metals are generally not toxic, but can cause a variety of water quality problems if present in large
quantities.
Heavy metals include arsenic, cadmium, chromium, lead, mercury, selenium, sulfur, and
several others. These metals vary significantly in their toxicity. For example, arsenic is quite
poisonous, but sulfur is not. Lead is toxic in large amounts, and in much lower amounts is thought
to cause developmental problems in small children. The toxicity of some metals depends on their
chemical state. Mercury is much more toxic in organic compounds than as a pure metal. Many of
us have chased little balls of mercury around after a thermometer broke, and suffered no ill effects,
while some organic mercury compounds are so toxic that one small drop on the skin can cause
almost instantaneous death. Hexavalent chromium (Cr
6+
) is extremely toxic, while trivalent
chromium (Cr
3+
) is much less so. (See the movie Erin Brockovich.)
Nutrients include nitrogen and phosphorus (and the metal potassium). Nitrogen is present as
nitrate (NO
3

-
) and nitrite (NO
2
-
). Nitrates can cause problems with drinking water, and phosphorus
can pollute surface waters.
Inorganic nonmetals include ammonia, bicarbonate, carbonate, chloride, cyanide, fluoride,
iodide, nitrite, nitrate, phosphate, sulfate, and sulfide. With the exception of cyanide, these are not
particularly toxic, but can contribute to low water quality if present in sufficient quantities.
Asbestos is an inorganic pollutant that is somewhat different from the others. Most toxic
substances are toxic due to their chemical activity. Asbestos is toxic, at least in air, because of its
physical properties. The small fibers of asbestos (a silicate mineral) can cause cancer when
breathed into the lungs. It has not been established whether it is toxic in drinking water.
© 2002 by CRC Press LLC
Radiologic parameters such as plutonium, radium, thorium, and uranium consist of both
natural materials and man-made products. These materials were produced in large quantities for
use in weapons and nuclear reactors, and many other uses (for example, lantern mantles and smoke
detectors). They can cause health hazards through either chemical or radiologic exposure. Some
radioactive materials, such as plutonium, are extremely toxic, while others such as uranium are less
so. High levels of radioactivity (alpha and beta particles and gamma rays) can cause acute health
problems, while long exposure to lower levels can lead to cancer. Radiologic parameters are
differentiated by isotope number, which varies for the same element depending on the number of
neutrons in the nucleus of each atom. For example, radium
224
and radium
226
have atomic weights of
224 and 226, respectively, but are the same element and participate in chemical reactions the same
way. Different isotopes have different levels of radioactivity and different half-lives (how long it
takes half of the material to undergo radioactive decay), so they are often tracked separately.

Inorganic pollutants in air include gaseous oxides such as carbon dioxide, sulfur dioxide, and
the oxides of nitrogen, which cause acid rain and may contribute to atmospheric warming (the
greenhouse effect). Chloride atoms in the atmosphere can damage the ozone layer, which protects
us from harmful ultraviolet radiation from the sun. Particulate matter is also significant, some of
which, for example, sea salt, is natural, but much of which is man-made. Colloidal-sized particles
formed by physical processes (dispersion aerosols) or chemical processes (condensation aerosols)
can cause smog and health problems if inhaled.
Organic parameters
Organic compounds are compounds that contain carbon, usually with hydrogen and often with
oxygen. Organics may contain other atoms as well, such as halides, nitrogen, and phosphorus. They
are usually segregated into volatile organic compounds (VOCs), and semivolatile organic
compounds (SVOCs). Hydrocarbons, chlorinated hydrocarbons, pesticides, and herbicides are also
organic compounds.
The delineation between volatiles and semivolatiles is not as easy as it sounds. SW-846, the
guidance document from EPA for analytical methods (EPA, 1980) describes volatile compounds as
“compounds which have boiling points below 200°C and that are insoluble or slightly soluble in
water.” Other references describe volatiles as those compounds that can be adequately analyzed by
a purge and trap procedure. Unfortunately, semivolatiles are described altogether differently. SW-
846 describes semivolatiles in their procedures as “neutral, acidic and basic organic compounds
that are soluble in methylene chloride and capable of being eluted without derivatization.” No
mention is made of the boiling points of semivolatile compounds, although it’s probably implicit
that their boiling points are higher than volatile compounds (Roy Widmann, pers. comm., 2001).
VOCs are organic compounds with a high vapor pressure, meaning that they evaporate easily.
Examples include benzene, toluene, ethylene, and xylene, (collectively referred to as BTEX),
acetone, carbon tetrachloride, chloroform, ethylene glycol, and various alcohols. Many of these
compounds are used as industrial solvents and cleaning fluids.
SVOCs are organic compounds with a low vapor pressure, so they resist evaporation. They
also have a higher boiling point than VOCs, greater than 200°C. Examples include anthracene,
dibenzofuran, fluorene (not to be confused with the halide fluorine), pentachlorophenol (PCP),
phenol, polycyclic aromatic compounds (PAHs), polychlorinated biphenyls (PCBs, Aroclor), and

pyrene. Some of these substances are used in manufacture of a wide variety of materials such as
plastics and medicine. Others are degradation products resulting from exposure of other organics to
the environment.
Halogenated compounds are organic compounds that have one or more of the hydrogens
replaced with a halide like fluorine, chlorine, or bromine. For example, 1,2-dibromoethane has the
first and second hydrogen replaced by bromine. One category of halogenated SVOCs,
polychlorinated biphenyls, were widely used in industry in applications such as cooling of
© 2002 by CRC Press LLC
transformers until banned by TSCA in 1976. They have high chemical, thermal, and biological
stability, which makes them very persistent in the environment.
Hydrocarbons consist of crude oil and various refined products. The have a wide range of
physical and chemical properties, and are widely dispersed in the environment. In some situations,
hydrocarbons are exempt from hazardous materials regulation. For example, crude oil is not
currently considered hazardous if spilled at the wellhead, but it is if spilled during transportation.
Chlorinated hydrocarbons can pose a significant health risk by contaminating drinking water
(Cheremisinoff, 2001). Three of these substances, trichlorethylene (TCE), trichloroethane (TCA),
and tetrachloroethylene (perchloroethylene or PCE), are widely used industrial solvents, and are
highly soluble in water, so a small quantity of material can contaminate a large volume of
groundwater.
Pesticides (insecticides) are widely distributed in the environment due to agricultural and
other uses, especially since World War II. Some pesticides such as nicotine, rotenone, and
pyrethrins are naturally occurring substances, are biodegradable, and pose little pollution risk.
Organochloride insecticides such as DDT, dieldrin, and endrin were widely used in the 1960s, but
are for the most part banned due to their toxicity and persistence in the food chain. These have
been largely replaced by organophosphates such as malathion, and carbamates such as carbaryl and
carbofuran.
Herbicides are also widely used in agriculture, and too often show up in drinking water. There
are several groups of herbicides, including bipyridilium compounds (diquat and paraquat),
heterocyclic nitrogen compounds (atrazine and metribuzin), chlorophenoxyls (2,4-D and 2,4,5-T),
substituted amides (propanil and alachlor), nitroanilines (trifluralin), and others. By-products from

the manufacture of pesticides and herbicides and the degradation products of these materials are
also significant problems in the environment.
Organic pollutants in the air can be a significant problem, including direct effects such as
cancer caused by inhalation of vinyl chloride, or the formation of secondary pollutants such as
photochemical smog.
Other parameters
There are a number of other parameters that may be tracked in an EDMS. Some are pollutants,
while others describe physical and chemical parameters that help better understand the site
geology, chemistry, or engineering. Examples include biologic parameters, field parameters,
geophysical measurements, operating parameters, and miscellaneous other parameters.
Biologic parameters (also called microbes, or pathogens if they are toxic) include fungi,
protozoa, bacteria, and viruses. Pathogens such as Cryptosporidium parvum and Giardia cause a
significant percentage (more than half) of waterborne disease. However, not all microbes are bad.
For example, bacteria such as Micrococcus, Pseudomonas, Mycobacterium, and Nocardia can
degrade hydrocarbons in the environment, both naturally and with human assistance, as a site
cleanup method.
Field parameters fall into two categories, The first are parameters measured at the time that
samples, such as water samples, are taken. These include pH, conductivity, turbidity, groundwater
elevation, and presence and thickness of sinkers and floaters. In some cases, such as field pH,
multiple observations may be taken for each sample, and this must be taken into consideration in
the design of the database system and reporting formats. The other category of field parameters is
items measured or observed without taking a sample. A variety of chemical and other
measurements can be taken in the field, especially for air monitoring, and increasingly for
groundwater monitoring, as sensors continue to improve.
Groundwater elevation is a special type of field observation. It can be observed with or
without sampling, and obtaining accurate and consistent water level elevations can be very
important in managing a groundwater project. This is because many other parameters react in
© 2002 by CRC Press LLC
various ways to the level of the water table. Issues such as the time of year, amount of recent
precipitation, tidal influences, and many other factors can influence groundwater elevation. The

EDMS should contain sufficient field parameter information to assist with interpretation of the
groundwater elevation data.
Geophysical measurements are generally used for site characterization, and can be done
either on the surface or in a borehole. For some projects this data may be stored in the EDMS,
while for others it is stored outside the database system in specialized geophysical software.
Examples of surface geophysical measurements include gravity, resistivity, spontaneous potential
(SP), and magnetotellurics. Borehole geophysical measurements include SP, resistivity, density,
sonic velocity, and radiologic parameters such as gamma ray and neutron surveys.
Operating parameters describe various aspects of facility operation that might have an
influence on environmental issues. Options for storage of this data are discussed in a previous
section, and the parameters include production volume, fluid levels, flow rates, and so on. In some
cases it is important to track this information along with the chemical data because permits and
other regulatory issues may correlate pollutant discharges to production volumes. Managing
operating parameters in the EDMS may require that the system be able to display calculated
parameters, such as calculating the volume of pollutant discharge by multiplying the concentration
times the effluent volume, as shown in Figure 54.
Miscellaneous other parameters include the RCRA characterization parameters (corrosivity,
ignitability, reactivity, and toxicity) as well as other parameters that might be measured in the field
or lab such as color, odor, total dissolved solids (TDS), total organic carbon (TOC), and total
suspended solids (TSS).
In addition, there are many other parameters that might be of importance for a specific project,
and any of these could be stored in the EDMS. The design of the EDMS must be flexible enough
to handle any parameter necessary for the various projects on which the software is used.
© 2002 by CRC Press LLC
CHAPTER 12
ENVIRONMENTAL LABORATORY ANALYSIS
Once the samples have been gathered, they are usually sent to the laboratory for analysis. How
well the laboratory does its work, from sample intake and tracking through the analysis and
reporting process, has a major impact on the quality of the resulting data. This chapter discusses
the procedures carried out by the laboratory, and some of the major laboratory analytical

techniques. A basic understanding of this information can be useful in working effectively with the
data that the laboratory generates.
The laboratory business is a tough business. In the current market, the amount of analytical
work is decreasing, meaning that the laboratories are having to compete more aggressively for what
work there is. They have cut their profit margins to the bone, but are still expected to promptly
provide quality results. Unfortunately this has caused some laboratories to close, and others to cut
corners to the point of criminal activities to make ends meet. Project managers must be ever
vigilant to make sure that the laboratories are doing their job with an adequate level of quality. The
good news is that there are still many laboratories performing quality work.
LABORATORY WORKFLOW
Because laboratories process a large volume of samples and must maintain a high level of
quality, the processes and procedures in the laboratory must be well defined and rigorously
followed. Each step must be documented in a permanent record. Steps in the process include
sample intake, sample tracking, preparation, analysis, reporting, and quality control.
Sample intake – The laboratory receives the samples from the shipper and updates the chain
of custody. The client is notified that the samples have arrived, the shipping container temperature
is noted, and the samples are scheduled for analysis.
Sample tracking – It is critical that samples and results be tracked throughout the processes
performed in the laboratory. Usually the laboratory uses specialized software called a laboratory
information management system or LIMS to assist with sample tracking and reporting.
Sample preparation – For most types of samples and analyses, the samples must be prepared
before they can be analyzed. This is discussed in the following section.
Analysis – Hundreds of different analytical techniques are available for analyzing different
parameters in various media. The most common techniques are described below.
Reporting – After the analyses have been performed, the results must be output in a format
suitable for use by the client and others. An electronic file created by the laboratory for delivery to
© 2002 by CRC Press LLC
the client is called an electronic data deliverable (EDD). Creation of the EDD should be done
using the LIMS, but too often there is a data reformatting step, or even worse a manual
transcription process, to get some or all of the data into the EDD, and these manipulation steps can

be prone to errors.
Quality control – The need to maintain quality permeates the laboratory process. Laboratory
quality control is discussed below, and aspects of it are also discussed in Chapter 15. Laboratory
QC procedures are determined by the level required for each site. For example, Superfund projects
generally require the highest level of QC and the most extensive documentation.
SAMPLE PREPARATION
Most samples must be prepared before they can be analyzed. The preparation process varies
depending on the sample matrix, the material to be analyzed, and the analytical method. The most
important processes include extraction and cleanup, digestion, leaching, dilution, and filtering.
Depending on the sample matrix, other procedures such as grinding and chemical manipulations
may be required.
Extraction and cleanup – Organic analytes are extracted to bring them into the appropriate
solvent prior to analysis (Patnaik, 1997). The extraction method varies depending on whether the
sample is liquid or solid. Extraction techniques for aqueous samples include liquid-liquid
(separatory funnel or continuous) and solid-phase. For solid samples, the methods include Soxhlett,
supercritical fluid, and sonication. Some extraction processes are repeated multiple times (such as
three) to improve the efficiency of extraction. Samples may undergo a cleanup process to improve
the analysis process and generate more reliable results. Cleanup methods include acid-base
partitioning, alumina column, silica gel, florisil, gel-permeation, sulfur, and permanganate-sulfuric
acid.
Digestion – Samples analyzed for metals are usually digested. The digestion process uses
strong acids and heat to increase the precision and accuracy of the measurement by providing a
homogeneous solution for analysis by removing metals adsorbed to particles and breaking down
metal complexes. Different digestion techniques are used depending on the analytical method and
target accuracy levels.
Leaching – Sometimes in addition to the concentration of a toxic substance in a sample, the
mobility of the substance is also of concern, especially for material headed for disposal in a
landfill. A leach test is used to determine this. Techniques used for this are the toxicity
characterization leaching procedure (TCLP), synthetic precipitate leaching procedure (SPLP),
and the EP toxicity test (EPTOX). In all three methods, fluids are passed through the solid material

such as soil, and the quantity of the toxic substance leached by the fluid is measured. TCLP uses a
highly buffered and mildly acidic aqueous fluid. In SPLP the fluid is slightly more acidic, varies by
geographic area (east or west of the Mississippi River), and is intended to more accurately
represent the properties of groundwater. EPTOX takes longer, tests for fewer parameters, and is no
longer widely used. The concentration of an analyte after leaching is not comparable to the total
concentration in the sample, so leached analyses should be so marked as such in the EDMS.
Dilution – Sometimes it is necessary to dilute the sample prior to analysis. Reasons for this
include that the concentration of the analyte may be outside the concentration range where the
analytical technique is linear, or other substances in the sample may interfere with the analysis
(matrix interference). A record of the dilution factor should be kept with the result. Dilution affects
the result itself as well as the detection limit for the result (Sara 1994, p. 11-11). For non-detected
results, the reported result based on the detection limit will be increased proportionately to the
dilution, and this needs to be considered in interpreting the results.
Filtering – The sample may or may not be filtered, either in the field or in the laboratory. If
the sample is not filtered, the resulting measurement is referred to as a total measurement, while if
it is filtered, it is considered a dissolved result. For filtered samples, the size of the openings in the
© 2002 by CRC Press LLC
filter (such as 1 micron) should be included with the result. Commonly, once a sample has been
filtered it is preserved. This information should also be noted with the result.
ANALYTICAL METHODS
Laboratories use many different methods to analyze for different constituents. Most of the
guidance on this comes from the EPA. The main reference for analytical methods in the
environmental field is the EPA’s SW-846 (EPA, 1980). SW-846 is the official compendium of
analytical and sampling methods that have been evaluated and approved for use in complying with
the RCRA regulations. It was first issued in 1980, and is updated regularly.
The tables in Appendix D show the recommended analytical methods for various parameters.
This section provides a general description of the methods themselves, starting with methods used
mostly for inorganic constituents, followed by methods for organic analysis. Additional
information on analytical methods can be found in many sources, including EPA (1980, 2000a),
Extoxnet (2001), Spectrum Labs (2001), SKC (2001), Cambridge Software (2001), NCDWQ

(2001), Scorecard.org (Environmental Defense, 2001), Manahan (2000, 2001), Patnaik (1997), and
Weiner (2000).
Inorganic methods
Many different methods are used for analysis of inorganic constituents. The most common
include titration, colorimetric, atomic absorption and emission spectrometry, ion-selective
electrodes, ion chromatography, transmission electron microscopy, gravimetry, nephelometric, and
radiochemical methods.
Titration is one of the oldest and most commonly used of the wet chemistry techniques. It is
used to measure hardness, acidity and alkalinity, chemical oxygen demand, non-metals such as
chlorine and chloride, iodide, cyanide, nitrogen and ammonia, sulfide and sulfite, and some metals
and metal ions such as calcium, magnesium, bromate, and bromide. Titration can be used on
wastewater, potable water, and aqueous extracts of soil and other materials. The method works by
slowly adding a standard solution of a known concentration to a solution of an unknown
concentration until the chemical reaction between the solutions stops, which occurs when the
analyte of concern in the second solution has fully reacted. Then the amount of the first solution
required to complete the reaction is used to calculate the concentration in the second. The
completion of the reaction is monitored using an indicator chemical such as phenolphthalein that
changes color when the reaction in complete, or with an electrode and meter. Titration methods
commonly used in environmental analyses include acid-base, redox, iodometric, argentometric, and
complexometric. Titration is relatively easy and quick to perform, but other techniques often have
lower detection limits, so are more useful for environmental analyses.
Colorimetric methods are also widely used in environmental analysis. Hardness, alkalinity,
chemical oxygen demand, cyanide, chloride, fluoride, ammonia, nitrite, nitrogen and ammonia,
phosphorus, phosphate and orthophosphate, silica, sulfate and sulfite, phenolics, most metals, and
ozone are among the parameters amenable to colorimetric analysis. For the most part, colorimetric
methods are fast and inexpensive. Aqueous substances absorb light at specific wavelengths
depending on their physical properties. The amount of monochromatic light absorbed is
proportional to the concentration, for relatively low concentrations, according to Beer’s law. First
the analyte is extracted, often into an organic solvent, and then a color-forming reagent is added.
Filtered light is passed through the solution, and the amount of light transmitted is measured using

a photometer. The result is compared to a calibration curve based on standard solutions to derive
the concentration of the analyte.
Atomic absorption spectrometry (AA) is a fast and accurate method for determining the
concentration of metals in solution. Aqueous and non-aqueous samples are first digested in nitric
© 2002 by CRC Press LLC

×