Chapter 6

Introduction Into the Environment

Abstract

When the use of organic chemical results in discharge into, and contamination of the environment, it becomes necessary to set standards for acceptable concentrations in water, air, soil, and biota. Monitoring of these concentrations in the environment and resultant biological effects, must be undertaken to ensure that the standards as set in any regulation are realistic and provide protection of the environment from all adverse effects. Furthermore, considerable attention continues to be focused on the regulation of the use of all organic chemicals and a primary aspect involves the prediction of the behavior and effects of a chemical, from its properties. With this concept the characteristics of the molecule govern the physicochemical properties of the compound which in turn influences transformation and distribution in the environment and the biological effects. This suggests that the transformation and distribution in the environment as well as any effects on the floral and faunal species can be predicted from the physicochemical properties of the chemical. However, the prediction of biological effects is the most complex of the set of predictions.

This chapter deals with the methods of introduction of organic chemicals into the environment and the means by which the chemicals may react with the environment.

Keywords

Dispersion; Dissolution; Emulsification; Evaporation; Leaching; Sedimentation; Adsorption; Spreading; Partitioning; Vapor pressure; Volatility; Water solubility; Total petroleum hydrocarbons

1 Introduction

When the use of organic chemical results in discharge into, and contamination of the environment, it becomes necessary to set standards for acceptable concentrations in water, air, soil, and biota (Chapter 8). Monitoring of these concentrations in the environment, and resultant biological effects, must be undertaken to ensure that the standards as set in any regulation are realistic and provide protection of the environment from all adverse effects. Furthermore, considerable attention continues to be focused on the regulation of the use of all organic chemicals and a primary aspect involves the prediction of the behavior and effects of a chemical, from its properties. With this concept the characteristics of the molecule govern the physicochemical properties of the compound which in turn influences transformation and distribution in the environment and the biological effects. This suggests that the transformation and distribution in the environment as well as any effects on the floral and faunal species can be predicted from the physicochemical properties of the chemical. However, the prediction of biological effects is the most complex of the set of predictions.

Industrial organic chemical manufacturers use and generate both large numbers and quantities of chemicals. In the past, the organic chemicals industry had introduced organic chemicals to all types of environmental ecosystems including air (through both fugitive emissions and direct emissions), water (direct discharge and runoff), and land (Table 6.1). The types of pollutants a single facility will release depend on (1) the type of process, (2) the process feedstocks, (3) the equipment in use, such as the reactor, and (4) the equipment and process maintenance practices, which can vary over short periods of time (such as from hour to hour) and can also vary with the part of the process that is underway. For example, for batch reactions in a closed vessel, the chemicals are more likely to be emitted at the beginning and end of a reaction step (that are associated with reactor or treatment vessel loading and product transfer operations) than during the reaction.

Table 6.1

Types of Releases From Industrial Processes

Release is an on-site discharge of a toxic chemical to the environment and include (1) emissions to the air, (2) discharges to bodies of water, (3) releases at the facility to land, as well as (4) the contained disposal into underground injection wells

Releases to air (point and fugitive air emissions): these releases include all air emissions from industry activity; point emissions occur through confined air streams as found in stacks, ducts, or pipes while fugitive emissions include losses from equipment leaks or evaporative losses from impoundments, spills, or leaks.

Releases to water (surface water discharges): these releases include any releases going directly to streams, rivers, lakes, oceans, or other bodies of water; any estimates for storm water runoff and nonpoint losses must also be included.

Releases to land: these releases include disposal of toxic chemicals in waste to on-site landfills, land treated or incorporation into soil, surface impoundments, spills, leaks, or waste piles; these activities must occur within the facility's boundaries for inclusion in this category.

Underground injection: this type of release is a contained release of a fluid into a subsurface well for the purpose of waste disposal.

Transfer is a transfer of toxic chemicals in wastes to a facility that is geographically or physically separate from the facility reporting under the toxic release inventory. The quantities reported represent a movement of the chemical away from the reporting facility and, except for off-site transfers for disposal, these quantities of chemicals do not necessarily represent entry of the chemicals into the environment

Transfers to publicly owned treatment works: include waste waters transferred through pipes or sewers to a publicly owned treatment works (POTW); treatment and chemical removal depend on the nature of the chemical and the treatment methods employed; chemicals that are not treated or destroyed by the publicly owned treatments works are generally released to surface waters or land filled within the sludge.

Transfers to recycling: these transfers include chemicals that are sent off-site for the purposes of regenerating or recovering still valuable materials; once these chemicals have been recycled, they may be returned to the originating facility or sold commercially.

Transfers to energy recovery: these transfers include wastes combusted off-site in industrial furnaces for energy recovery; treatment of a chemical by incineration is not considered to be energy recovery.

Transfers to treatment: these transfers are wastes moved off-site for either neutralization, incineration, biological destruction, or physical separation; in some cases, the chemicals are not destroyed but are prepared for further waste management.

Transfers to disposal are wastes taken to another facility for disposal generally as a release to land or as an injection underground.

t0010

The organic chemical pollutants that are most likely to present ecological risks are those that are (1) highly bioaccumulative, building up to high levels in floral and faunal tissues even when concentrations in the ecosystem remain relatively low, and (2) highly toxic, so that they cause harm at comparatively low doses. In addition, atmosphere-water interactions that control the input and outgassing of persistent organic pollutants (POPs) in aquatic systems are critically important in determining the life cycle and residence times of these compounds and the extent of contamination. Although the effects of various types of organic chemical pollutants are usually evaluated independently, many ecosystems are subject to multiple pollutants, and their fate and impacts are intertwined. For example, the effects of nutrient deposition in an ecosystem can alter the methods by which the organic contaminants are assimilated, bioaccumulated, and the means by which the organisms in the ecosystem are affected.

Of all the organic chemical pollutants released into the environment by anthropogenic activity, POPs are among the most dangerous and need extreme measures for removal (Chapter 1). POPs are chemical substances that persist in the environment, bioaccumulate through the food web, and pose a risk of causing adverse effects to human health and the environment. This group of priority pollutants consists of pesticides (such as DDT), industrial chemicals (such as polychlorinated biphenyls, PCBs), and unintentional by-products of industrial processes (such as dioxins and furans). POPs can be transported across international boundaries far from their sources, even to regions where they have never been used or produced. Consequently, POPs pose a threat to the environment and to human health, all on a global scale.

Releases into the environment of organic chemicals that persist in the ecosystem (rather than undergo some form of biodegradation) lead to an exposure level that is not only subject to the length of time the chemical remains in circulation (in the environment) but also on the number of times that the organic chemical is recirculated before it is ultimately removed from the ecosystem. Typically, POPs are pesticides, industrial chemicals, or unwanted by-products of industrial processes that have been used and disposed for decades (prior to the inception of the various regulations and often without due regard for the environment) but have more recently been found to share a number of significant characteristics that need consideration before disposal is planned, including: (1) persistence in the environment insofar as these organic chemicals resist degradation in air, water, and sediments, (2) bioaccumulation insofar as these organic chemicals accumulate in living tissues at concentrations higher than those in the surrounding environment, and (3) long-range transport insofar as these organic chemicals can travel great distances from the source of release through air, water, and the internal organs of migratory animals, any of which can result in the contamination of ecosystems thousands of significant distances (up to thousands of miles) away from the source of the chemicals.

Briefly and by way of explanation, bioaccumulation is a process by which persistent environmental pollution leads to the uptake and accumulation of one or more contaminants, by organisms in an ecosystem. The amount of a pollutant available for exposure depends on its persistence and the potential for its bioaccumulation. Any chemical (including an organic chemical) is considered to be capable of bioaccumulation if the chemical has a degradation half-life in excess of 30 days or if the chemical has a bioconcentration factor (BCF) > 1000 if the log Kow is > 4.2:

BCF=concentrationinbiota/concentrationinecosystem

si1_e

The BCF indicates the degree to which a chemical may accumulate in fish (and other aquatic animals, such as mussels, etc.) by transport across the gills or other membranes, excluding feeding. Bioconcentration is distinct from food-chain transport, bioaccumulation, or biomagnification. The BCF is a constant of proportionality between the chemical concentration in flora or fauna in an ecosystem. It is possible, for many organic chemicals, to estimate the BCF from the octanol-water partition coefficients (Kow) (Bergen et al., 1993):

Logbioconcentrationfactor=mlogKow+b

si2_e

In terms of actual numbers, for many lipophilic organic chemicals, the BCF can be calculated using the regression equation:

logBCF=2.3+0.76×logKow

si3_e

Furthermore, empirical relationships between the octanol-water partition coefficients and the BCF can be developed on a chemical-by-chemical basis.

On this note, it is worth defining the source of chemical contaminant insofar as chemical contaminants can originate from a point source or a nonpoint source (NPS). Point source of pollution is a single identifiable source of pollution which may have negligible extent, distinguishing it from other pollution source geometries. On the other hand, NPS pollution generally results from land runoff, precipitation, atmospheric deposition, drainage, seepage, or hydrologic modification. NPS pollution, unlike pollution from industrial and sewage treatment plants, comes from many diffuse sources and is often caused by rainfall or snowmelt moving over and through the ground. As the runoff moves, it picks up and carries away natural and human-made pollutants, finally depositing them into lakes, rivers, wetlands, coastal waters, and groundwater.

Typically, POPs are highly toxic and long-lasting (hence the name persistent) and cause a wide range of adverse effects to environmental flora and fauna, including disease and birth defects in humans and animals—some of the severe human health impacts of POPs include (1) the onset of cancer, (2) damage to the central nervous system, (3) damage to the peripheral nervous system, (4) damage to the reproductive system, and (5) disruption of the immune system. Moreover, POPs do not respect international borders, the serious environmental and human health hazards created by these chemicals affect not only developing countries, where systems and technology for monitoring, tracking, and disposing them can be weak or nonexistent, but also affect developed countries. As long as the chemical can be transported by air, water, and land, no country is immune from the effects of these chemicals.

In the last four decades it has become increasingly clear that the chemical and allied industries, such as the pharmaceutical industry, has been faced with serious environmental problems. Many of the classical organic synthetic processes have broad scope but often generate extreme amounts of chemical waste. As a result, the chemical industry, the pharmaceutical industry, the various fossil fuel industries, as well as any other allied chemical industries have been subjected to increasing environmental regulation pressure to minimize or, preferably, eliminate any form of chemical waste. An illustrative example is provided by the manufacture of phloroglucinol (1,3,5-trihydroxybenzene, C6H6O3), a reprographic chemical and pharmaceutical intermediate. Until the latter part of the 20th century, phloroglucinol was produced mainly from 2,4,6-trinitrotoluene (TNT):

u06-01-9780128044926
1,3,5-Trihydroxybenzene (phloroglucinol)
u06-02-9780128044926
2,4,6-Trinitrotoluene (TNT)

Unfortunately, as a result of this process, for every kilogram of phloroglucinol produced, approximately 40 kg of solid waste were generated and also the waste contained such environmental nasties as chromium sulfate [Cr2(SO4)3], ammonium chloride (NH4Cl), ferrous chloride (FeCl2), and potassium bisulfate (KHSO4). This process was eventually discontinued as the costs associated with the disposal of this chromium-containing waste approached or exceeded the selling price of the product. This decision to terminate the process, seemingly based on process economics, boded well for the environment.

Indeed, an analysis of the amount of waste formed in processes for the manufacture of a range of fine chemicals and pharmaceutical intermediates has revealed that the generation of tens of kilograms of waste per kilogram of desired product was not exceptional in the organic chemical industry. This led to the introduction of the E (environmental) factor (kilograms of waste per kilogram of product) as a measure of the environmental footprint of manufacturing processes in various segments of the chemical industry. Thus

E=kilogramsofwaste/kilogramofproduct

si4_e

The factor can be conveniently calculated from a knowledge of the number of tons of raw materials purchased and the number of tons of product sold, the calculation being for a particular product or a production site or even a whole company. A higher E factor means more waste and, consequently, a larger environmental footprint—thus, the ideal E factor for any process is 0.

However, in the context of environmental protection and to be all inclusive, the E factor is the total mass of raw materials plus ancillary process requirements minus the total mass of product, all divided by the total mass of product. Thus, the E factor should represent the actual amount of waste produced in the process, defined as everything but the desired product and takes the chemical yield into account and includes reagents, solvent losses, process aids, and (in principle) even the fuel necessary for the process. Water has been generally excluded from the E factor as the inclusion of all process water could lead to exceptionally high E factors in many cases and made meaningful comparisons of the technical factors (E factors excluding water use) for processes difficult. However, in the modern environmentally-conscious era, there is no reason for water requirement to be omitted since the disposal of process water is an environmental issue. Moreover, use of the E factor has been widely adopted by many of the chemical industries and the pharmaceutical industries in particular. Thus, a major aspect of process development recognized by process chemists and by process engineers is the need for determining an E factor—whether or not it is called by that name (i.e., E factor) but chemicals in vs. chemical out has become a major yardstick in many industries.

It is clear that the E factor increases substantially on going from bulk chemicals to fine chemicals and then to pharmaceuticals. This is partly a reflection of the increasing complexity of the products, necessitating not only processes which use multistep syntheses, but is also a result of the widespread use of stoichiometric amounts of the reagents (Chapters 2 and 3), i.e., the required amounts of the reagents (some observers would advocate the stoichiometric amounts of the reagents plus 10%) to accomplish conversion of the starting organic chemical to the product(s). A reduction in the number of steps of a process for the synthesis of organic chemicals will, in most cases (but not always), lead to a reduction in the amounts of reagents and solvents used and hence a reduction in the amount of waste generated. This has led to the introduction of the concepts of step economy and function oriented synthesis (FOS) of pharmaceuticals. The main issues behind the concept of FOS is that the structure of an active compound, which may be a natural product, can be reduced to simpler structures designed for ease of synthesis while retaining or enhancing the biological activity. This approach can provide practical access to new (designed) structures with novel activities while at the same time allowing for a relatively straightforward synthesis.

As noted above, a knowledge of the stoichiometric equation allows the process chemist or process engineer to predict the theoretical minimum amount of waste that can be expected. This led to the concept of atom economy or atom utilization to quickly assess the environmental acceptability of alternatives to a particular product before any experiment is performed. It is a theoretical number, that is, it assumes a chemical yield of 100% and exact stoichiometric amounts and disregards substances which do not appear in the stoichiometric equation. In short, the key to minimizing waste is precision or selectivity in organic synthesis which is a measure of how efficiently a synthesis is performed. The standard definition of selectivity is the yield of product divided by the amount of substrate converted, expressed as a percentage.

Organic chemists distinguish between different categories of selectivity: (1) chemoselectivity, which relates to competition between different functional groups and (2) regioselectivity, which is the selective formation of one regioisomer, for example, ortho vs. para substitution in aromatic ring systems. However, one category of selectivity was, traditionally, largely ignored by organic chemists: the atom selectivity or atom utilization or atom economy. The virtually complete disregard of this important parameter by chemists and engineers has been a major cause of the waste problem in the manufacture of organic chemicals. Quantification of the waste generated in chemicals manufacturing, by way of E factors, served to illustrate the omissions related to the production of chemical waste and focus the attention of fine chemical companies, the pharmaceutical companies, and the petrochemical companies on the need for a paradigm shift from a concept of process efficiency, which was exclusively based on chemical yield, to a need that more focused on the elimination of waste chemicals and maximization of raw materials utilization.

Many of the global environmental changes forced by human activities are mediated through the chemistry of the environment (Andrews et al., 1996; Schwarzenbach et al., 2003; Manahan, 2010; Spellman, 2016). Important changes include the global spread of air pollution, groundwater pollution, and pollution of the oceans which increase the concentration of tropospheric oxidants (including ozone), stratospheric ozone depletion, and global warming (the so-called greenhouse effect). Since the onset and establishment of the agricultural revolution and the industrial revolution, the delicate balance between physical, chemical, and biological processes within the Earth system has been perturbed as a result. Example of the causes of perturbation of the Earth systems include (1) the exponential growth in the world population, (2) the use of increasing amounts of fossil fuel, (3) fossil fuel-related emissions of carbon to the atmosphere, and (4) the intensification of agricultural practices including the more frequent use of fertilizers. The observed increase in the atmospheric abundance of carbon dioxide (CO2) has been ascribed (correctly or incorrectly) mainly to fossil fuel burning (Speight and Islam, 2016), although biomass destruction is an important secondary source of carbon dioxide emissions. Atmospheric concentrations are additionally influenced by exchanges of carbon with the ocean and the continental biosphere (Firor, 1990; WMO, 1992; Calvert, 1994; Goody, 1995).

The progressive modification and fertilization of the terrestrial biosphere are believed to have caused the observed increase in atmospheric nitrous oxide (N2O), a tropospheric greenhouse gas (GHG) and a source of reactive species in the stratosphere. Methane (CH4), which also contributes to greenhouse forcing (the technical term for the influence of the greenhouse effect that causes a shift in the climate)—this can occur due to changes in the level of gasses that share two properties: they are transparent to visible light, but absorb the infrared, which we typically perceive as heat, and also plays an important role in the photochemistry of the troposphere and the stratosphere, is produced by biosphere-related processes (wetlands, livestock, landfills, biomass burning) as well as by leakage from gas distribution systems in various countries (Calvert, 1994). The global atmospheric concentration of methane has also grown in the past. Observed increases in the abundance of tropospheric ozone (O3), which contribute to deteriorating air quality, result from complex photochemical processes involving industrial and biological emissions of nitrogen oxides, hydrocarbons, and certain other organic compounds. Ozone is a strong absorber of solar ultraviolet radiation and also contributes to greenhouse forcing. Anthropogenic emissions of sulfur resulting from combustion of sulfur-containing fuels and also from coal combustion without the necessary end-of-pipe gas cleaning protocols, coal burning in highly populated and industrialized regions of the Northern Hemisphere, and the related increase in the aerosol load of the troposphere, have contributed to regional pollution and have probably produced a cooling of the surface in these regions by backscattering a fraction of the incoming solar energy.

Finally, the bad news is that the rapid increase in the abundance of industrially manufactured chlorofluorocarbons in the atmosphere produced an observed depletion in stratospheric ozone and the formation each spring (since the late 1970s) of an ozone hole over Antarctica. By way of explanation, the ozone hole is not technically a hole where no ozone is present, but it is actually a region of exceptionally depleted ozone in the stratosphere over the Antarctic that happens at the beginning of Southern Hemisphere spring (August–October). The good news is that, as a result of environmental caution and a reduction in terms of the release of contaminants, the ozone hole has diminished in size (i.e., the depletion of ozone has stopped) and may even be on a turnaround to an increase in the amount of ozone in the area.

Although toxic and hazardous chemicals are produced by the various chemicals industries, frequent reference is made here to the chemical products produced by the crude oil refining industry, without any attempt to point to this industry as the major polluter. The chemicals produced in the various cured oil-derived products offer a wide range of properties and behavior that makes the crude oil-derived products suitable for this text.

In summary, many of the specific chemicals in petroleum are hazardous because of their chemical reactivity, fire hazard, toxicity, and other properties. In fact, a simple definition of a hazardous chemical (or hazardous waste) is that it is a chemical substance (or chemical waste) that has been inadvertently released, discarded, abandoned, neglected, or designated as a waste material and has the potential to be detrimental to the environment. Alternatively, a hazardous chemical may be a chemical that may interact with other (chemical) substances to give a product that is hazardous to the environment.

2 Release Into the Environment

For the purposes of this text, it is assumed that any organic chemicals released into the environment are hazardous chemicals following from the lists of chemicals produced by the US Environmental Protection Agency as well as environmental agencies in many other countries (Appendix). Thus, it is not only safe, but necessary, to assume that any organic chemical (except chemicals that are indigenous to the ecosystem into which they exist but in quantities close to the indigenous amounts) have the potential to be hazardous to the environment.

Contamination by chemicals is a global issue and there is no single company that should shoulder all of the blame. Past laws and regulations (or the lack thereof) allowed unmitigated disposal of chemicals and discharge of chemicals into the environment. These companies were not breaking the law; it is a matter of there being insufficient laws (the fault of various level of government) that protected the environment. As a result, toxic chemicals are found practically in all ecosystems on Earth, thus affecting biodiversity, agricultural production, and water resources. At the end of the various chemical life cycles, chemicals are recycled or disposed as part of waste. The inappropriate management of such waste results in negative impacts on the environment.

As already stated, of all the pollutants released into the environment by human activity, POPs among the most dangerous to environmental flora and fauna are pesticides, various industrial chemicals, or unwanted by-products of industrial processes that have been used for decades but more recently been found to share a number of disturbing characteristics, including: (1) persistence, which means that the chemicals resist degradation in air, water, and sediments, (2) bioaccumulation, which means that the chemicals accumulate in living tissues at concentrations higher than those in the surrounding environment, and (3) long-range transport, which means that the chemicals can travel a considerable distance from the source of release through air, water, and migratory animals, often contaminating areas miles away from any known source. On the environmental side, POPs are highly toxic and long-lasting, and cause an array of adverse effects on flora and fauna.

Thus, many organic chemicals have toxic, carcinogenic, mutagenic, or teratogenic (causing developmental malformations) effects on environmental flora and fauna and are designated either as Acutely Hazardous Waste or Toxic Waste by the United States Environmental Protection Agency (https://www.epa.gov/hw). Substances found to be fatal to humans in low doses or, in the absence of data on human toxicity, have been shown to have an oral LD50 toxicity (lethal dose at 50% concentration) of < 2 mg L− 1, or a dermal LD50 of < 200 mg kg− 1 or is otherwise capable of causing or significantly contributing to an increase in serious irreversible, or incapacitating reversible illness are designated as Acute Hazardous Waste (https://www.epa.gov/hw). Materials containing any of the toxic constituents so listed are to be considered hazardous waste, unless, after considering the following factors it can reasonably be concluded (by the Department of Environmental Health and Safety) that (1) the waste is not capable of posing a substantial present or potential hazard to public health or (2) the waste is not capable of posing a substantial present or potential hazard to the environment when improperly treated, stored, transported or disposed of, or otherwise managed.

Briefly, within this text and for environmental purposes, chemicals are subdivided into two classes: (1) organic chemicals and (2) inorganic chemicals. Furthermore, classification occurs insofar as organic chemicals are classified as volatile organic compounds (VOCs) or semivolatile organic compounds (on occasion, the word chemicals is substituted for the word compounds without affecting the definition). The first class of organic compounds, the VOCs, is subdivided into regulated compounds and unregulated compounds. Regulated compounds have maximum contaminant levels, while unregulated compounds do not. Regulated compounds generally (but not always) have low boiling points, or low boiling ranges, and some are gases. Many of these chemicals can be detected at extremely low levels by a variety of instrumentation, including the human nose! In the case of organic chemicals, sources for VOCs typically are petroleum refineries, fuel stations, naphtha (i.e., dry cleaning solvents, paint thinners, cleaning solvents for auto parts) and, in some cases, refrigerants that are manufactured from petrochemicals.

The constituents of the second class of organic compounds, the semivolatile compounds, typically have high boiling points, or high boiling ranges, and are not always easily detected by the instrumentation that may be used to detect the VOCs (including the human nose). Some of the common sources of contamination are high boiling petroleum products (such as automotive lubricating oils and machinery lubricating oils), pesticides, herbicides, fungicides, wood preservatives, and a variety of other chemicals that can be linked to the organic chemicals industry.

Regulations are in place that set the maximum contamination concentration levels that are designed to ensure environmental and public safety. There are primary and secondary standards for inorganic chemicals. Primary standards are for those chemicals that cause neurological damage, cancer, or blood disorders. Secondary standards are developed for other environmental reasons. In some instances, the primary standards are referred to as the Inorganic Chemical Group. The secondary standards are referred to as the General Mineral Group and General Physical Testing Group.

However, despite the nature of the environmental regulations and the precautions taken by the organic chemicals industry, the accidental release of nonhazardous organic chemicals and hazardous organic chemicals into the environment has occurred and, without being unduly pessimistic, will continue to occur (by all industries—not wishing to select any particular industry as the only industry that suffers accidental release of organic chemicals into the environment). It is a situation that, to paraphrase chaos theory, no matter how well the preparation, the unexpected is always inevitable. It is, at this point that the environmental scientist and engineer has to identity (through careful analysis) the nature of the chemicals and their potential effects on the ecosystem(s) (Smith, 1999). Although petroleum itself and its various products are complex mixtures of many organic chemicals (Chapters 2 and 3), the predominance of one particular chemical or one particular class of chemicals may offer the environmental scientist or engineer an opportunity for predictability of behavior of the chemical(s).

Thus, when a spill of organic chemicals occurs the primary processes determining the fate of organic chemicals are (1) dispersion, (2) dissolution, (3) emulsification, (4) evaporation, (5) leaching, (6) sedimentation, (7) spreading, and (8) wind. These processes are influenced by the physical properties of the organic chemicals (especially if the chemicals are constituents of a mixture), spill characteristics, environmental conditions, and physicochemical properties of the spilled material after it has undergone any form of chemical transformation (Chapter 7).

2.1 Dispersion

For the purposes of this text, the term dispersion encompasses all phenomena which give rise to the proliferation of substances through the man-made and natural environment. Thus, the physical transport of oil droplets into the water column is referred to as dispersion. This is often a result of water surface turbulence, but also may result from the application of chemical agents (dispersants). These droplets may remain in the water column or coalesce with other droplets and gain enough buoyancy to resurface. Dispersed oil tends to biodegrade and dissolve more rapidly than floating slicks because of high surface area relative to volume. Most of this process occurs from about half an hour to half a day after the spill.

Emissions of many chemicals of concern occur into the air initially, from where they are dispersed into other media. In fact, air is one of the main carriers of chemical carcinogens to humans (Corvalán and Kjellström, 1996). Many chemicals emitted into the air, for instance from combustion processes, tend to become associated with particulate matter. Removal from the air occurs through a range of complex processes involving photo-degradation, and particle sedimentation and/or precipitation (known respectively as dry and wet deposition). Volatile organic chemicals and semivolatile organic chemicals may undergo several cycles of evaporation and precipitation, which can also make chemicals more accessible to photochemical or biodegradation.

The dispersion of chemicals released into the environment has been the focus of much attention because of the realization that the dispersion behavior of organic chemicals can be markedly different when different chemicals are considered. Accidents that involve organic chemicals give rise to a new class of problems in dispersion prediction for the following reasons: (1) the material is, in almost all cases, stored as a liquid, which may appear as a gas after a spill, (2) the modes of release can vary widely and geometry of the source can take many forms and the initial momentum of the spill may be significant, and (3) in some cases, a chemical transformation also takes place as a result of reaction with water vapor in the ambient atmosphere.

In addition, the physical properties of the organic chemical(s) usually result in interactions with the surrounding ecosystem, especially if the chemical is a reactive liquid and has the potential of interacting with the air, water, or soil. This reactivity will influence the dispersibility of the chemical. Moreover, if the release occurs over a short time-scale, compared to the steady-state releases characteristic of many chemical releases problems, this can give rise to the complication of predicting dispersion for time-varying releases and to uncertainty in individual predictions resulting from variability about the behavior of a mixture. Also, the dispersing chemical, which is typically denser that air, may form a low-level cloud that is sensitive to the effects of man-made and natural obstructions and of topography.

Wind (Aeolian) transport (relocation by wind) can also occur and is particularly relevant when dust from organic carbonaceous solids (such as coke dust) and catalyst dust are considered. Dust becomes airborne when winds traversing arid land with little vegetation cover pick up small particles such as catalyst dust, coke dust, and other refinery debris and send them skyward after which the movement of pollutants in the atmosphere is caused by transport, dispersion, and deposition. Dispersion results from local turbulence, that is, motions that last less than the time used to average the transport. Deposition processes, including precipitation, scavenging, and sedimentation, cause downward movement of pollutants in the atmosphere, which ultimately remove the pollutants to the ground surface.

2.2 Dissolution

The solubility (dissolution) characteristics of organic molecules in water are complex and very much dependent upon the structure and properties of the organic chemicals(s). Solubility is the property of organic chemicals which might be a gas, a liquid, or a solid (the solute) that dissolves in a solvent which might also be a gas, a liquid, or a solid organic chemical (the solvent). The solubility of a gas, a liquid, or a solid organic chemical depends on the physical and chemical properties of the solute and solvent as well as on temperature, pressure, and the pH of the solution. The extent of the solubility of a substance in a specific solvent is measured as the saturation concentration, where adding more solute does not increase the concentration of the solution and begins to precipitate the excess amount of solute. In terms of environmental organic chemistry, the solvent is typically a liquid, which can be a pure organic chemical (such as hexane or toluene) or a mixture of organic chemicals (such as naphtha or kerosene).

The extent of solubility of the solute ranges widely, from infinitely soluble (without limit) (i.e., the solute is fully miscible with the solvent, such as ethanol in water) to poorly soluble (such as some polynuclear aromatic systems in water)—the term insoluble is often applied to poorly or very poorly soluble compounds and (in the world of organic chemistry) a common threshold to describe an organic chemical as insoluble is a solubility < 0.1 g per 100 mL of solvent. However, solubility of an organic compound in a solvent should not be confused with the ability to of a solvent to dissolve a solute, because the solution might also occur because of a chemical reaction (reactive solubility).

Solubility of a solute in a solvent applies not only to environmental organic chemistry but to areas of chemistry, such as (alphabetically): biochemistry, geochemistry, inorganic chemistry, physical chemistry, and organic chemistry. In all cases, solubility depends on the physical conditions (temperature, pressure, and concentration of the solute) and the enthalpy and entropy directly relating to the solvents and solutes concerned. By far the most common solvent in chemistry is water which is a solvent for most ionic compounds as well as a wide range of organic chemicals. This is a crucial factor in the chemical phenomena known as acidity and alkalinity as well as in much of the area known as environmental organic chemistry.

In addition, the term dissolved organic carbon (DOC) is used as a broad classification for organic molecules of varied origin and composition within aquatic systems (the aquasphere) and the dissolved fraction of organic carbon is an operational classification. The source of the DOC in marine systems and in freshwater systems depends on the body of water. In general, organic chemical compounds are a result of decomposition processes from dead organic matter such as plants or marine organisms. When water contacts highly organic soils, these components can drain into rivers and lakes as DOC. Whatever the source of the DOC, it is also extremely important in the transport of metals in aquatic systems—certain metals can form extremely strong organo-metallic complexes with DOC which enhances the solubility of the metal in aqueous systems while also reducing the bioavailability of the metal.

In terms of organic chemicals that are not naturally occurring, knowledge of the structure and properties can be used to examine relationships between the solubility properties of an organic chemical and its structure, and vice versa. In fact, structure dictates function which means that by knowing the structure of an organic chemical it may be (but not always) possible to predict the properties of the chemical such as its solubility, acidity or basicity, stability, and reactivity. In the context of environmental distribution of a chemical, predicting the solubility of an organic molecule is a useful component of knowledge.

At the molecular level, solubility is controlled by the energy balance of the intermolecular forces between solute-solute, solvent-solvent, and solute-solvent molecules. But without getting too fay into a study of such forces, the simple, very useful, and practical empirical rule that is quite reliable which is like dissolves like which is based on the polarity of the organic chemical systems insofar as polar organic chemicals typically dissolve in polar solvents (such as water, alcohols) and nonpolar organic chemicals typically dissolve in nonpolar solvents (such as nonpolar hydrocarbon solvents). The polarity of organic molecules is determined by the presence of polar bonds due to the presence of electronegative atoms (such as nitrogen and oxygen) in polar functional groups such as amine derivatives (RNH2) and alcohol derivatives (ROH) (Chapter 2). Furthermore, since the polarity of an organic chemical is related to the presence of a functional group, the solubility characteristics of an unknown organic contaminant can provide evidence (as well as evidence form other analytical techniques) (Chapter 5) for the presence (or absence) of an organic functional group:

SolventSolubility or Complete Miscibility
WaterAlcohols, amines, acids, selected (but not all) esters, ketones, and aldehyde derivatives
5% aqueous NaHCO3aCarboxylic acid derivatives
5% aqueous NaOHaCarboxylic acid derivatives and phenol derivatives
5% aqueous HClaAmine derivatives
Diethyl etherMost organic chemicals

a A 5% w/w solution of the chemical in water.

Most organic molecules are relatively nonpolar and are usually soluble in organic solvents (such as diethyl ether, dichloromethane, chloroform, and hexane) but not in polar solvents (such as water). If the organic chemical is soluble in water, this denotes that the chemical is polar and therefore soluble in water which denotes a high ratio of polar group(s) to the nonpolar hydrocarbon chain, such as a low molecular weight organic chemical that contains a hydroxy function (glyph_sbndOH), or an amino function (glyph_sbndNH2) or a carboxylic acid function (glyph_sbndCO2H) group, or a higher molecular weight organic chemical that contains two or more functional (polar) groups. The presence of an acidic carboxylic acid or basic amino group in a water-soluble compound can be detected by measurement of the pH of the solution—a low pH (pH ≤ 7) indicates the presence of an acidic function while a high pH (pH ≥ 7) indicates the presence of a basic function. Thus, organic chemicals that are insoluble in water can become soluble in an aqueous environment if they form an ionic species when treated with an acid or a base—the ionic form (for example the sedum salt of a carbocyclic acid, RCO2 Na+) is much more polar than the carboxylic acid.

The solubility of carboxylic acid derivatives and phenol derivatives in sodium aqueous hydroxide is due to the formation of the polar (ionic) carboxylate ions (glyph_sbndCO2) or the polar (ionic) phenoxide (C6H5O) ions since they are much stronger acids than water, and therefore the acid-base equilibria lie far to the right of the equation:

RCO2H+OHRCO2+H2O

si5_e

ArOH+OHArO+H2O

si6_e

Carboxylic acid derivatives, but not phenol derivatives, are also stronger acids than carbonic acid (H2CO3), and are therefore also soluble in aqueous sodium bicarbonate (NaHCO3 solution):

RCO2H+HCO3RCO2+H2O+CO2

si7_e

The solubility of amine derivatives in dilute aqueous acid is in accordance with the amine derivatives being stronger bases than water, and the amine derivatives are converted to the polar ammonium ion by protonation (through reaction with a proton):

RNH2+H3O+RNH3++H2O

si8_e

Amine derivatives are the only common class of organic compounds which are protonated in dilute aqueous acid.

Thus, dissolution is the solubilization of organic chemicals in water. Many of the acutely toxic components of oils such as benzene, toluene, and xylene will also dissolve into water but only to a minute extent. Nevertheless, when the volume of water (as in a lake or in running water such as a river) is large the transportation of the benzene, toluene, ethylbenzene, and the xylene isomers (BTEX) through dissolution can be substantial. This process also occurs quickly after a discharge of the chemical(s) but tends to be less important than evaporation. For example, in a typical discharge into the aquasphere, generally < 5–10% v/v of the benzene is lost to dissolution while > 90–95% v/v can be lost to evaporation. The polynuclear aromatic compounds (Chapter 2) offer a different scenario.

For alkylated polynuclear aromatic compounds, solubility is inversely proportional to the number of rings and extent of alkylation (i.e., the number of alkyl moieties attached to the polynuclear aromatic ring system). The dissolution process is thought to be much more important in rivers because natural containment within the river system (i.e., the water, the character of the river banks, and the river bed) may prevent spreading, thereby reducing the surface area of the chemical slick (the chemicals on the surface of the water) and thus retard evaporation. However, river turbulence must not be ignored since turbulence increases the potential for mixing and dissolution.

Groundwater that is contaminated with organic chemicals tends to be enriched in aromatic derivatives relative to other organic chemicals constituents. Relatively insoluble hydrocarbons may be entrained in water through adsorption on to clay (kaolinite) particles suspended in the water or as an agglomeration of oil droplets (microemulsion) in the water. In cases where groundwater contains only dissolved hydrocarbons, it may not be possible to identify the original organic chemicals because only a portion of the free product will be present in the dissolved phase. If a hydrocarbon mixture has been spilled into the aquasphere, initially the whole mixture floats on top of the groundwater but any constituents with a tendency to water-solubility will be extracted from the mixture. Groundwater containing entrained hydrocarbons will have a gas chromatographic fingerprint that is a combination of the total mixture plus enhanced amounts of the soluble aromatics minus losses to the water. Generally, dissolved aromatics may be found quite far from the origin of a spill but entrained hydrocarbons may be found in water close to the organic chemical source. Oxygenates, such as methyl-t-butyl ether, are even more water soluble than aromatics and are highly mobile in the environment.

2.3 Emulsification

An emulsion is a mixture of two or more liquids that are usually immiscible. Examples include crude oil and water which can form an oil-in-water emulsion, wherein the oil is the dispersed phase, and water is the dispersion medium. In addition, crude oil and water can also form a water-in-oil emulsion, wherein water is the dispersed phase and crude oil is the external phase. Whether an emulsion of oil and water exists as a water-in-oil emulsion or an oil-in-water emulsion depends on the volume fraction of both phases and the type of emulsifier (surfactant) present. Multiple emulsions are also possible, including a water-in-oil-in-water emulsion and an oil-in-water-in-oil emulsion. Emulsions, being liquids, do not exhibit a static internal structure—the droplets dispersed in the liquid matrix (the dispersion medium) are usually assumed to be statistically distributed.

The stability of an emulsion refers to the ability of the emulsion to resist change in its properties over time. There are three types of instability in emulsions: (1) flocculation, (2) creaming, and (3) coalescence. Flocculation occurs when there is an attractive force between the droplets, so they form flocs. Coalescence occurs when droplets bump into each other and combine to form a larger droplet, so the average droplet size increases over time. Creaming occurs when the droplets rise to the top of the emulsion under the influence of buoyancy. Use of a surface-active agent (surfactant) can increase the stability of an emulsion so that the size of the droplets does not change significantly with time and the emulsion is defined as stable. Most emulsions contain droplets with a mean diameter of more than around 1 μm (1 micron, 1 × 10− 6 m) however mini-emulsions and nano-emulsions can be formed with droplet sizes in the 100–500 nm (nanometer, 1 × 10− 9 m) range, and with proper formulation, highly stable microemulsion can be prepared having droplets as small as a few nanometers.

Crude oil and crude oil products as well as many other organic chemicals can form water-in-oil emulsions (where water is incorporated into oil) or in the form of a mousse as weathering occurs. This process is significant because, for example, the apparent volume of the organic chemical may increase dramatically, and the emulsification will slow the other weathering processes, especially evaporation of the volatile organic chemicals.

Emulsification in the aquasphere (especially in marine environment) depends primarily on the composition of the crude oil (or the crude oil product) and the turbulent regime of the water mass. The most stable emulsions such as water-in-oil contain from 30% to 80% v/v water and usually appear after strong storms in the zones of spills of heavy oil or the higher boiling crude oil product that have an increased content of nonvolatile fractions (especially asphaltene constituents) (Speight, 2014). These types of emulsions can exist in the marine environment for over several months (100 days or more) in the form of (what is colloquially called) a chocolate mousse (which is not recommended for human consumption!). The stability of these emulsions usually increases (demulsification) with decreasing temperature—the reverse emulsions, such as oil-in-water emulsion (where droplets of oil are suspended in water) are much less stable because surface-tension forces quickly decrease the dispersion of oil.

This demulsification process can be slowed with the help of emulsifiers which are surface-active substances with strong hydrophilic properties that are used to eliminate the prolonged effects of spills of crude oil and crude oil products. Emulsifiers help to stabilize oil emulsions and promote dispersing oil to form microscopic (invisible) droplets which accelerates the decomposition of oil products in the water.

2.4 Evaporation

Evaporation (the opposite of condensation) occurs when a liquid organic chemical becomes a gas (or vapor), sublimation is the phenomenon that occurs when a solid becomes a gas without the conversion of the solid to liquid and thence to a gas.

Evaporation:Liquidgas

si9_e

Sublimation:Solidgas

si10_e

Both phenomena are important aspects of the environmental organic chemical cycle since both processes involve disappearance of organic contaminant for the water or land but the contaminant does appear in the atmosphere. Of the two processes, the evaporation is the most common process since organic chemicals that go through the sublimation processes are not as obvious or as common as chemicals that evaporate.

Many factors affect the evaporation process. For example, if the air is already saturated with other chemicals (or the humidity is high) there is typically little chance of a liquid evaporating as quickly as when the air is not saturated (or humidity is low). In addition, the air pressure also affects the evaporation process since, under high air pressure, an organic chemical is much more difficult to evaporate. Temperature also affects the ability of an organic chemical to evaporate.

Evaporative processes are very important in the weathering of volatile organic chemical products, and may be the dominant weathering process for naphtha, gasoline, and other mixtures that contain low-boiling organic chemicals. Automotive gasoline, aviation gasoline, and some grades of jet fuel (e.g., JP-4) contain 20–99% highly volatile constituents (i.e., constituents with less than nine carbon atoms). The evaporative processes typically begin immediately after a volatile organic chemical is discharged into the environment. For spilled crude oil, the amount lost to evaporation can typically range from approximately 20% to 60% v/v. Some low-boiling petroleum-derived products such as one-ring and two-ring aromatic hydrocarbons and/or low molecular weight alkane derivatives (having < 15 carbon atoms) may evaporate entirely. In fact, a substantial fraction of higher molecular weight organic chemicals may also evaporate.

The primary factors that control evaporation are the composition of the oil, slick thickness, temperature and solar radiation, wind speed, and wave height. While evaporation rates increase with temperature, this process is not restricted to warm climates. For the Exxon Valdez incident, which occurred in cold conditions (Mar. 1989), it has been estimated that appreciable evaporation occurred even before all the oil escaped from the ship, and that evaporation ultimately accounted for 20% v/v of the crude oil. Most of this process occurs within the first few days after the spill. However, it is not unusual for the evaporation process to be active simultaneously with other processes to remove any volatile organic chemicals.

2.5 Leaching

Leaching is a natural process by which water-soluble substances (such as water-soluble organic chemicals or hydrophilic organic chemicals) are washed out from soil or waste disposal areas (such as a landfill). These leached out chemicals (leachates) can cause pollution of surface waters (ponds, lakes, rivers, and the sea) and subsurface water (groundwater aquifers). Thus, leaching is the process by which (in the current context) organic contaminants are released from the solid phase into the water phase under the influence of dissolution, desorption, or complexation processes as affected by acidity or alkalinity (the pH value). The process itself is universal, as any material exposed to contact with water will leach components from its surface or its interior depending on the porosity of the material under consideration. Leaching often occurs naturally with soil contaminants such as organic chemicals with the result that the chemicals end up in potable waters.

Many organic chemicals occur in a mixture of different components in a solid (such as the carbonaceous deposits on petroleum coke). In order to separate the desired solute constituent or remove an undesirable solute component from the solid phase, the solid is brought into contact with a liquid during which time the solid and liquid are in contact and the solute or solutes can diffuse from the solid into the solvent, resulting in separation of the components originally in the solid (leaching, liquid-solid leaching). In addition, leaching may also be referred to as extraction because the solute is being extracted from the solid or the proceeds may also be referred to as washing because a contaminant is removed from a solid with water.

Leaching is affected by (1) soil texture, (2) structure, and (3) water content of the soil. For example, in terms of soil texture, the proportions of sand, silt, and clay affect the movement of water through the soil. Coarse-textured soil containing more sand particles have large pores and is highly permeable which allow the water to move rapidly through the soil and, in fact, organic chemicals (such as pesticides) carried by water through coarse-textured soil are more likely to reach and contaminate groundwater. On the other hand, clay-textured soils have low permeability and tend to retain more water and adsorb more organic chemicals from the water. This slows the downward movement of chemicals, helps increase the chance of degradation and adsorption to soil particles, and reduces the chance of groundwater contamination.

In terms of soil structure, loosely packed soil particles allow speedy movement of water through the soil while tightly compacted soil holds water back and does not allow the water to move freely through it. Plant roots penetrate soil, creating excellent water channels when they die and rot away. These openings and channels may permit relatively rapid water movement even through clay-containing soil. On the other hand, the amount of water already in the soil has a direct bearing on whether rain or irrigation results in the recharging of groundwater and possible leaching of organic chemicals into an aquifer. Soluble chemicals are more likely to reach groundwater when the water content of the soil approaches or is at saturation. Saturation is typical in the spring when rain and snowmelt occurs but when soil is dry the added water just fills the pores in the soil near the soil surface, making it unlikely that the water will reach the groundwater supply.

Leaching processes introduce hydrocarbon into the water phase by solubility and entrainment. Leaching processes of organic chemical products in soils can have a variety of potential scenarios. Part of the aromatic fraction of a spill of organic chemicals on to soil may partition into water that has been in contact with the contamination.

2.6 Sedimentation or Adsorption

Sedimentation occurs when particles in suspension settle out of the fluid in which they are entrained and come to rest against a barrier, which is typically the basement of a waterway. Settling is due to the motion of the particles through the fluid in response to the forces acting on them, which in an ecosystem, can be due to gravity. The term sedimentation is often used as the opposite of erosion, i.e., the terminal end of sediment transport. Settling occurs when suspended particles fall through the liquid, whereas sedimentation is the termination of the settling process.

Sedimentation can be generally classified into three different types that are all applicable to organic chemicals. Type 1 sedimentation is characterized by particles that settle discretely at a constant settling velocity and typically these particles settle as individual particles and do not flocculate or stick to others during settling. Type 2 sedimentation is characterized by particles that flocculate during sedimentation and because of this the particle size is constantly changing and therefore their settling velocity keeps changing. Type 3 sedimentation (also known as zone sedimentation) involves particles that are at a high concentration (e.g., > 1000 mg L− 1) such that the particles tend to settle as a mass and a distinct clear zone and sludge zone are present. Zone settling occurs in active sludge sedimentation and sedimentation of sludge thickeners.

Sediments typically consist of a few weight percent of organic matter and the balance being shared by various mineral constituents. The organic carbon environment acts as an attractor or accumulator of hydrophobic compounds, such as polychlorobiphenyl derivatives (PCBs). However, organic carbon in sediment comes in different forms that may have very different sorption capacities for hydrophobic organic compounds. In addition to natural sources such as vegetative debris, decayed remains of plants and animals, and humic matter, sediment organic carbon is also derived from particles of coal, coke, charcoal, and soot that are known to have extremely high sorption capacities.

Many organic chemicals (especially crude oil and crude oil-derived products) are less dense than water or are buoyant in water. However, in areas with high levels of suspended sediment, organic chemicals may be transported to the river, lake, or ocean floor through the process of sedimentation. The organic chemicals may adsorb on to sediments and sink—most of this process occurs from about 2 to 7 days after the spill. Furthermore, hydrophobic organic compounds, such as polynuclear aromatic hydrocarbons (PNAs) and PCBs, bind strongly to sediments and can serve as a long-term source of contaminants in water bodies and biota long after the original source has been removed. The inherent heterogeneity of most sediments makes it difficult to describe in terms of bulk sediment physicochemical parameters such as total organic carbon content, surface area, and particle size distribution. Therefore, management of sediments and the control of sediment contaminants are among the most challenging and complex problems faced by the environmental scientists and the environmental engineers and will become increasingly more so as other organic contaminants enter the environment in greater quantities.

In terms of the sedimentation of organic chemicals (Crompton, 2012), some of the chemicals may be adsorbed on the suspended material (especially if the suspended material is clay or another highly adsorptive mineral) and deposited to the bottom. This mainly happens in shallow waters where particulate matter is abundant and water is subjected to intense mixing, usually through turbulence. Simultaneously, the process of biosedimentation can also occur when plankton and other organisms absorb the organic chemical. The suspended forms of the organic chemicals undergo intense chemical and biological (microbial in particular) decomposition in the water. However, this situation radically changes when the suspended organic chemical reaches the lake bed, river bed, or sea bed and the decomposition rate of the organic chemical(s) buried on the bottom abruptly drops. The oxidation processes slow down, especially under anaerobic conditions in the bottom environment and the organic chemical(s) accumulated inside the sediments can be preserved for many months and even years.

2.7 Spreading

The movement of organic chemical contaminants through the subsurface is complex and is difficult to predict since different types of contaminants react differently with soils, sediments, and other geologic materials and commonly travel along different flow paths and at different velocity. Contaminants released at the source areas (typically on the surface) infiltrate into the subsurface and migrate downward by gravity through the vadose zone (the zone that extends from the top of the ground surface to the water table). When low-permeability soil units are encountered, the contaminants can also spread laterally along the permeability contrast.

Most organic chemical contaminants are introduced into the subsurface by percolation through soil strata and the interactions between the soil and a chemical contaminant are important for assessing the fate and transport of the contaminant in the subsurface, especially in the groundwater system. Contaminants that are highly soluble, such as salts of carboxylic acid derivatives can move readily from surface soil to saturated materials below the water table and often occurs during and after rainfall events. Those contaminants that are not highly soluble may have considerably longer residence times in the surface strata (the soil zone). Some contaminants adsorb readily onto soil particles and slowly dissolve during precipitation events, resulting in migration into groundwater—this is typical of the mode of transport for chemicals such as trichloroethylene (CCl2glyph_dbndCHCl). Liquids spilled onto surface soils can migrate downward or can evaporate, which limits their potential for reaching the water table. Once below the water table, organic chemical contaminants are also subject to dispersion (mechanical mixing with uncontaminated water) and diffusion (dilution by concentration gradients).

Many organic contaminants (such as crude oil and crude oil-derived products) begin to spread immediately after entry into the environment. The viscosity of the oil, its pour point, and the ambient temperature will determine how rapidly the oil will spread, but light oils typically spread more rapidly than heavy oils. The rate of spreading and ultimate thickness of the oil slick will affect the rates of the other weathering processes. For example, discharges that occur in geographically contained areas (such as a pond or slow-moving stream) will evaporate more slowly than if the oil is allowed to spread.

3 Types of Chemicals

The ability to collect and preserve a sample that is representative of the site is a critically important step in identification of chemicals in the environment (Dean, 1998; Patnaik, 2004; Speight, 2005). Obtaining representative environmental samples is always a challenge due to the heterogeneity of different sample matrices. Additional difficulties are encountered with petroleum hydrocarbons due to the wide range in volatility, solubility, biodegradation, and adsorption potential of individual constituents and the procedures used for sample collection and preparation must be legally defensible.

As described above, organic chemicals can enter the air, water, and soil when they are produced, used, or disposed. The impact of these chemicals on the environment is determined by the amount of the chemical that is released, the type and concentration of the chemical, and where it is found. Some chemicals can be harmful if released to the environment even when there is not an immediate, visible impact. On the other hand, some chemicals are of concern as they can work their way into the food chain and accumulate and/or persist in the environment for many years.

On a beneficial note, various types of organic chemicals (including mixtures such as crude oil and various crude oil-derived products) are biodegraded faster than many individual and more complex chemicals (Appendix, Tables A2–A6). Nevertheless, different types of organic chemicals and crude oils (as well as the crude oil products even though given similar names) are biodegraded at different rates in the same ecosystem. A crude oil product is a complex mixture of organic chemicals and contains within it less persistent constituents and more persistent constituents. The range between these two extremes is often extensive since the different organic chemicals have different physical and chemical properties and estimating the behavior and fate of any organic chemical (or even a particular crude oil-derived product) is very difficult.

The relative proportion of hazardous constituents present in any collection of organic chemicals (crude oil-derived products included) is variable and rarely consistent because of site differences. Therefore, the extent of the contamination will vary from one site to another and, in addition, the farther a contaminant progresses from low molecular weight to high molecular weight the greater the occurrence of polynuclear aromatic hydrocarbons, complex ring systems (not necessarily aromatic ring systems), as well as an increase in the composition of the semivolatile organic chemicals or the nonvolatile organic chemicals. These latter organic chemical constituents (many of which are not so immediately toxic as the volatiles) can result in long-term/chronic impacts to the flora and fauna of the environment. Thus, any complex mixture of organic chemicals should be analyzed for the semivolatile compounds which may pose the greatest long-term risk to the environment.

In addition to large spills of organic chemicals, crude oil-based hydrocarbons are released into the environment from natural seeps as well as NPS urban runoffs. Acute impacts from massive one-time spills are obvious and substantial but the impact of organic chemicals from small spills and chronic releases are the subject of much speculation. Clearly, such inputs of chemicals have the potential for significant environmental impact, but the effects of chronic low-level discharges can be minimized by the net assimilative capacities of many ecosystems, resulting in little detectable environmental harm.

Short-term (acute) hazards of lighter, more volatile, and water-soluble aromatic compounds (such as benzenes, toluene, ethylbenzene, and the xylene isomers, BTEX) include potential acute toxicity to aquatic life (especially in relatively confined areas) as well as potential inhalation hazards to other faunal species (including humans). However, the compounds which pass through the water column often tend to do so in small concentrations and/or for short periods of time, and fish and other pelagic species or generally mobile species can often swim away to avoid impacts from spilled oil in open waters.

Briefly, pelagic species are species that frequent the pelagic zone which is any water in a sea or lake that is neither close to the bottom nor near to the shore. The pelagic zone can be described in terms of an imaginary water column that extends from the surface of the sea almost to the sea bed. Conditions differ deeper in the water column such that as pressure increases with depth, the temperature drops and less light penetrates. Depending on the depth, the water column, rather like the atmosphere of the Earth (Chapter 1) may be divided into different layers. Pelagic life decreases with increasing depth and is affected by light intensity, pressure, temperature, salinity, the supply of dissolved oxygen, the supply of nutrients, and the submarine topography (bathymetry). In deep water, the pelagic zone (sometimes referred to as the open-ocean zone) can be contrasted with water that is near to the coast or water above the continental shelf.

Long-term (chronic) potential hazards of lighter, more volatile, and water-soluble aromatic compounds include contamination of groundwater. Chronic effects of benzene, toluene, ethylbenzene, and the xylene isomers include changes in the liver and harmful effects on the kidneys, heart, lungs, and nervous system. At the initial stages of a release of these organic chemicals, when the benzene-derived compounds are present at their highest concentrations, acute toxic effects are more common sooner rather than later. These noncarcinogenic effects include subtle changes in detoxifying enzymes and liver damage. Generally, the relative aquatic acute toxicity of petroleum will be the result of the fractional toxicities of the different hydrocarbons present in the aqueous phase. There are also indications that naphthalene-derived chemicals have a similar effect.

Organic chemicals are weathered according to the individual physical properties and chemical properties, but during this process living species within the local environment may be affected via one or more routes of exposure, including ingestion, inhalation, dermal contact, and, to a much lesser extent, bioconcentration through the food chain. Aromatic compounds of concern include alkylbenzene derivatives, toluene derivatives, naphthalene derivatives, and polynuclear aromatic hydrocarbons (PNAs). Moreover, the impact on both, atmosphere and hydrosphere must be assessed when considering the implications from a release of chemical containing significant quantities of these single-ring aromatic compounds.

By way of explanation, bioconcentration is the accumulation of a chemical in or on an organism and is also the process by which a chemical concentration in an organism exceeds that in the surrounding environment as a result of exposure of the organism to the chemical. Bioconcentration can be measured and assessed and these include: (1) octanol-water partition coefficient (Kow), (2) the bioconcentration factor, BCF, (3) the bioaccumulation factor, BAF, and (4) the biota-sediment accumulation factor, BSAF. Each of these factors can be calculated using either empirical data or measurements as well as from mathematical models (Mackay, 1982). The BCF can also be expressed as the ratio of the concentration of a chemical in an organism to the concentration of the chemical in the surrounding environment and is a measure of the extent of chemical sharing between an organism and the surrounding environment. Thus

BCF=concentrationinbiota/concentrationinecosystem

si1_e

The BCF can also be related to the octanol-water partition coefficient (Kow) which is correlated with the potential for a chemical to bioaccumulate in floral and faunal organisms. The BCF can be predicted from the octanol-water partition coefficient (Bergen et al., 1993):

Logbioconcentrationfactor=mlogKow+b

si12_e

Kow=concentrationinoctanol/concentrationinwater=Co/Cwatequilibrium

si13_e

4 Physical Properties and Distribution in the Environment

To effectively monitor changes in the environmental behavior of organic chemicals that are of most concern, it is extremely important to understand how these chemicals typically behave in natural systems and, in particular, in specific ecosystems. Equally important is an understanding of how these chemicals might respond to specific best management practices. Some of the discharged organic chemicals only become problems at high concentrations that impair the beneficial uses of these ecosystems. An effective monitoring program explicitly considers how these chemicals may change as they move from a source into the groundwater, surface water, or into the soil. This includes an understanding of how a specific organic chemical may be introduced or mobilized within an ecosystem, how the chemical moves through an ecosystem, and the transformations that may occur during this process (Chapter 7).

However, the entry of chemicals into the environment and the distribution of these chemicals within the environment is often complex and there have been many occasions when a significant amount of a chemical (or a mixture of chemicals) has entered an ecosystem and the effects of contamination are well defined. Generally, the assumption is that the organic chemical (or a mixture thereof) does not rapidly diffuse away, but remains in the immediate vicinity at a noticeably high concentration or perhaps moves, but in such a way that concentration levels of the chemical remain high as it moves. Such cases would normally occur when large quantities of a substance were being stored, transported, or otherwise handled in concentrated form.

Thus, due to leakages, spills, improper disposal, and accidents during transport, organic compounds have become subsurface contaminants that threaten various resources, the most important of which are drinking water resources.

4.1 Chemical and Biochemical Properties

One strategy to remediate such polluted environments, especially the subsurface, is to make use of the degradative capacity of bacteria. It is often sufficient to supply the subsurface with nutrients containing elements such as nitrogen and phosphorus. However, anaerobic processes have advantages such as low biomass production and good electron acceptor availability, and they are sometimes the only possible solution for cleanup and protection of the environment.

Whereas hydrocarbons are oxidized and completely mineralized under anaerobic conditions in the presence of electron acceptors such as nitrate (NO3), iron (Fe), sulfate (SO4), and carbon dioxide (CO2), chlorinated organic compounds and nitroaromatic compounds are reductively transformed to products that are not always benign. For the often persistent polychlorinated compounds, reductive dechlorination leads to harmless products or to compounds that are aerobically degradable. The nitroaromatic compounds (ArNO2) are first reductively transformed to the corresponding amines (ArNH2) and can subsequently be bound to the humic fraction of the soil in an aerobic process—humic constituents of soil are the major organic constituents of soil that are produced by biodegradation of dead organic matter.

Such new findings and developments give hope that in the near future contaminated aquifers can efficiently be remediated, a prerequisite for a sustainable use of the precious subsurface drinking water resources.

4.2 Partitioning

The increasing awareness of chemicals in the environment, their disposition, and their ultimate fate has created the need to find reliable mechanisms to assess the environmental behavior and effects of new chemicals. Whether or not a chemical will pose a hazard to the environment will depend on the concentration levels it will reach in various ecosystems and whether or not those concentrations are toxic to the floral and faunal species within the ecosystem. It is, therefore, important to determine expected environmental distribution patterns of chemicals in order to identify which areas will be of primary environmental concern (McCall et al., 1983).

In order to assess the behavior of an organic chemical in the environment, typical physical and chemical properties play a role but a property that is not often considered in the partitioning of the organic chemical in an ecosystem is a function of the chemical and physical structure of the chemical. This leads to a determination of the way in which the chemical or the mixture of chemicals are distributed among the different environmental phases (McCall et al., 1983). These phases may include air, water, organic matter, mineral solids, and even organisms.

More specifically, the focus of partitioning studies is on the equilibrium distribution of an organic chemical that is established between the phases the associated issue problem of calculating the distribution of a compound between the different phases (partitioning equilibrium). There are many situations in which it is correct to assume that phase transfer processes are fast compared to the other processes, such as chemical transformation of the organic contaminant to other (benign or hazardous) chemicals (Chapter 7) that play a role in determining the fate of contaminants. In such cases, it is appropriate to describe the distribution of the chemical as a change of phase by the equilibrium approach using the valid assumption that an equilibrium condition will be reached with the chemical passing to one phase or remaining distributed between different phases. By way of explanation, an example of phase equilibrium is the partitioning of a contaminant (such as carboxylic acid or a carboxylic acid derivative—that is, the sodium salt of the acid) between the pore water and other water in the bed of a sediment and solids in sediment beds. Thus, from the equilibrium partition coefficient it is possible to calculate the rate of transfer of an organic chemical contaminant across interfaces and the rate at which such transfer can be anticipated to occur.

However, before too much excitement is generated at the thought of the answers that a study of partitioning will provide, it must be recognized that there are many situations where an equilibrium of the chemical between the phases is not reached. However, some observers (justifiably) find the information useful insofar as the data are used (correctly or incorrectly) to characterize what the equilibrium distribution of the chemical would be if sufficient time (often difficult to define) is allowed. However, in such cases a quantitative description of the potential partitioning equilibrium can be employed to estimate the possible direction of the transport of the organic chemical contaminant from one environmental ecosystem to another. If this is the case, it may be possible to evaluate whether or not the chemical component(s) of a contaminant (such as a solvent or naphtha) will continue to dissolve in the groundwater and/or volatilize into the overlying soil. Both options are viable, but (hopefully) the data will provide the most likely option to occur, given the position of the chemical within the underground chemical and geological system. Thus, the goal of any partitioning study of the ability of an organic contaminant is to gain insight into the role played by the chemical (and physical) structure of a contaminant in determining the fate of the contaminant in the environment (McCall et al., 1983).

However, when dealing with phase partitioning parameters, it is necessary to develop an understanding of the intramolecular interactions and the intermolecular interactions (Chapter 5) between the organic chemical(s) and the specific molecular environment in which the chemical is spilled or disposed. Thus, there is the necessity to understand the means by which structural groups function within the contaminants and the means by which the functional groups are related to chemical and physical behavior (Chapters 2 and 5). Therefore, there must be an effort to understand (1) the interactions arising from contacts of functional groups within a molecule, i.e., the intramolecular interactions, (2) the influence of these intramolecular interactions on the existence of the molecule in the environment, (3) the interactions arising from contacts of functional groups with the functional groups of another molecule, i.e., the intermolecular interactions, and (4) the influence of these intermolecular interactions on the existence of the molecule in the environment. However, partition coefficients are not the only means of estimating the transfer of organic chemicals between phases. It is also valuable to correlate the partition coefficients on the basis of solubility of the chemical(s) which are convenient and readily understood measurable expressions of single-phase activity coefficients (Cole and Mackay, 2000).

Thus, this section presents an examination of the pertinent compound properties and environmental factors that are needed for quantifying such partitioning.

4.2.1 Acid-Base Partitioning

Acid-base partitioning (also called pH partitioning) is the tendency for acids to accumulate in basic fluid compartments and bases to accumulate in acidic regions. The reason for this phenomenon is that acids become negatively electric charged in basic fluids, since they donate a proton. On the other hand, bases become positively electric charged in acidic liquids because the base receives a proton (H+). Since electric charge decrease the membrane permeability of substances, once an acid enters a basic fluid and becomes electrically charged, then it cannot escape that compartment with ease and therefore accumulates, and vice versa with bases. Thus, by manipulating the pH of the aqueous layer, the partitioning of a solute can be changed.

4.2.2 Air-Water Partitioning

The air-water partition coefficient (Kaw) is the constant of proportionality between the concentration of a chemical in air and its concentration in water at low partial pressures and below its saturation limits in either air or water. The coefficient can be estimated from Henry's law constant and is sometimes referred to as the dimensionless.

It is essential to have reliable data for the air-water partition coefficient or Henry's law constant for the compounds under investigation for elucidating the environmental dynamics of many natural and anthropogenic compounds. When a compound (here referred to as the solute) is introduced into the environment, it tends to diffuse from phase to phase in the direction towards establishing equilibrium between all phases. Frequently, the physical-chemical properties of the solute dictate that it will partition predominantly into a different phase from the one into which it is normally emitted. For example, benzene emitted in waste water will tend to partition or transfer from that water into the atmosphere where it becomes subject to atmospheric photolytic degradation processes. A knowledge of the air-water partition characteristics of a solute is thus important for elucidating where the solute will tend to accumulate and also in calculating the rates of transfer between the phases. Conventionally these rates are expressed as the product of a kinetic constant such as a mass transfer coefficient (or diffusivity divided by a diffusion path length) and the degree of departure from equilibrium which exists between the two phases. Elucidating the direction and rate of transfer of such solutes thus requires accurate values for the Henry's law.

The transfer between the atmosphere and bodies of water is one of the key processes affecting the transport of many organic compounds in the environment. For neutral compounds, at dilute solution concentration in pure water, the air-water distribution ratio is referred to as the Henry's Law constant (KH or Kaw). For real aqueous solutions (i.e., solutions that contain many other chemical species), the term air-water distribution ratio is often used which, for practical purposes, is approximated by the Henry's Law constant. The Henry's Law constant KH can be approximated as the ratio of a compound's abundance in the gas phase to that in the aqueous phase at equilibrium.

Xaq=Xg

si14_e

K=Pi/CwatmLmol1

si15_e

In this equation, Pi is the partial pressure of the gas phase of the chemical and Cw is the molar concentration of the chemical in water.

Knowledge of the air-water partitioning behavior of VOCs and GHGs is important in a number of environmental applications. For example, VOCs emitted from open process streams in a paper mill can promote ground level ozone and lead to respiratory problems in humans. Therefore, it is important to have reliable estimates of the amount of VOCs in the atmosphere in contact with open process streams. In global climate models, the partitioning of carbon dioxide and methane between the atmosphere and ocean water is of current interest since oceans represent a substantial storage reservoir for these gases.

4.2.3 Molecular Partitioning

Partition coefficients are the ratio of the concentration of an organic compound in two phases that are in equilibrium with each other. For example, in a two-layer system of water (bottom layer) and an organic solvent (top layer), an organic compound will be in one or the other of the layers. After stirring and allowing time for the layers to settle, the organic compound could well be in both phases, albeit to a different extent (concentration, C) in each phase. The partition coefficient (K) is

K=Corganic/Cwater

si16_e

From this equation, a high value of K suggests that the compound is not very water soluble but is more soluble in the organic solvent that is if the organic compound is lipophilic (or hydrophobic).

Molecular partitioning occurs when an organic compound dissolves in each of two immiscible solvent phases and is measured by the partition coefficient or distribution coefficient which is the ratio of concentrations of the compound in a mixture of the two immiscible phases at equilibrium. This ratio is therefore a measure of the difference in solubility of the compound in these two phases. Most commonly, one of the solvents is water while the second is hydrophobic such as 1-octanol (1-hydroxy octane, CH3CH2CH2CH2CH2CH2CH2CH2OH). Hence the partition coefficient measures how hydrophilic or hydrophobic a chemical substance is.

The hydrophobic nature (hydrophobicity) of a compound can give an indication of the relative ease that a compound might be taken up in groundwater to pollute waterways. The partition coefficient can also be used to predict the mobility of species in groundwater and, in the field of hydrogeology, the octanol-water partition coefficient (Kow) is used to predict and model the migration of dissolved hydrophobic organic compounds in soil and groundwater.

4.2.4 Octanol-Water Partitioning

The distribution of nonpolar organic compounds between water and natural solids (e.g., soils, sediments, and suspended particles) or organisms, can be viewed in many cases as a partitioning process between the aqueous phase and the bulk organic matter present in natural solids or in biota. More recently, environmental chemists have found similar correlations with soil humus and other naturally occurring organic phases. These correlations exist because the same molecular forces controlling the distribution of compounds between water-immiscible organic solvents and water also determine environmental partitioning from water into natural organic phases.

The octanol-water partition coefficient is a key parameter in understanding and predicting the environmental fate and transport behavior of organic chemicals (Lyman et al., 1990; Boethling and Mackay, 2000). Kow is often used as a surrogate for the lipophilicity of a chemical and its tendency to concentrate in organic phases such as within plant lipids or fish from the aqueous solution. Chemicals with relatively low Kow values are considered relatively hydrophilic and will tend to have high water-solubility and low BCFs (Lyman et al., 1990). It is also used in the prediction of other parameters including water solubility and the organic carbon-water partition coefficient.

Thus, in order to simulate lipids (fats) in organic media (biota), n-octanol (CH3CH2CH2CH2CH2CH2CH2CH2OH) was selected as model compound for partitioning experiments. Thus, the partition coefficient that best describes lipophilicity is the octanol-water partition coefficient (Kow). The values of the octanol-water partition coefficient typically fall into the range 102–107 and it is often more convenient to use the common logarithm of Kow. The octanol-water partition coefficient is defined simply by

Kow=Coctanol/Cwater

si17_e

Coctanol is the molar concentration of the organic compound in the octanol phase and Cwater is the molar concentration of the organic compound in water when the system is at equilibrium.

For a series of neutral nonpolar compounds partitioning between octanol and water, the Kow value is determined largely by the magnitude of the aqueous activity coefficient (a measure of the dissimilarity between the organic solute and the aqueous solvent). In other words, the major factor that determines the magnitude of the partition constant of a nonpolar or moderately polar organic compound between an organic solvent and water is the incompatibility of the compound with water. The nature of the organic solvent is generally of secondary importance.

4.2.5 Sorption Partitioning

The persistence of organic pollutants in topsoil, their migration to groundwater, and the evaluation of the degree of contamination expected in a groundwater system after an accidental spill or as consequence of the presence of a waste disposal site, are problems of particular environmental concern which require the knowledge of the sorption characteristics of the pollutants to be investigated as well as the knowledge of the type of soil and its characteristics. Sorption also affects volatility of organic pollutants, their bioavailability and bioactivity, phytotoxicity, and chemical or microbial transformations (Delle Site, 2001).

The sorption of an organic chemical on a natural solid is a very complicated process which involves many sorbent properties, besides the physicochemical properties of the chemical itself. These properties are especially reflective of the relative amount of the mineral and organic material in soil/sediment and their respective composition with associated physical characteristics. Also, different regions of a soil or sediment matrix may contain different types, amounts, and distributions of surfaces and of soil organic material, even at the particle scale.

The extent to which an organic chemical partitions itself between solid and solution phases is determined by several physical and chemical properties of both the chemical and the soil or sediment aqueous solutions. However, in most cases, the tendency of a chemical to be adsorbed or desorbed can be expressed in terms of the organic carbon partition coefficient (KOC) which is largely independent of soil or sediment properties (Lyman et al., 1982). Thus, the organic carbon partition coefficient is a chemical specific adsorption parameter and may be determined as the ratio of the amount of chemical adsorbed per unit weight of organic carbon in the soil or sediment to the concentration of the chemical in solution at equilibrium:

KOC=microgramadsorbedpergramoforganiccarbon/microgrampermLofsolution

si18_e

Factors which affect measured values of the organic carbon partition coefficient include (1) temperature, (2) acidity or alkalinity, measures as the pH, (3) salinity, (4) the concentration of dissolved oxygen, (5) suspended particulates, and (6) the solids-to-solution ratio. The presence of other chemicals in a complex mixture could alter the activity coefficient of the chemical in the water, the pH of the water, or the solubility of the chemical in water, and, consequently, the sorption of the chemical to soils and sediments. The degree of adsorption will not only affect chemical mobility but will also affect volatilization, photolysis, hydrolysis, and biodegradation. Adsorption of organic chemicals will also occur on minerals free of organic matter. It may be significant under certain conditions such as: (1) clay minerals with a very large surface area, (2) situations where cation exchange occurs, such as for dissociated organic bases, (3) situations where clay-colloid-induced polymerization occurs, and/or (4) situations where chemisorption is a factor (Lyman et al., 1982).

The distribution of an organic solute between sorbent and solvent phases results from its relative affinity for each phase, which in turn relates to the nature of forces which exist between molecules of the solute and those of the solvent and sorbent phases. The type of interaction depends on the nature of the sorbent as well as the physicochemical features of the sorbate ~ hydrophobic or polar at various degrees.

The physical sorption processes involve interactions between dipole moments—permanent or induced—of the sorbate and sorbent molecules. The relatively weak bonding forces associated with physical sorption are often amplified in the case of hydrophobic molecules by substantial thermodynamic gradients for repulsion from the solution in which they are dissolved. Chemical interactions involve covalent bond and hydrogen bond. Finally, electrostatic interactions involve ion-ion and ion-dipole forces. In a more detailed way, the type of interactions and the approximate values of energy associated are (1) van der Waals interactions, (2) hydrophobic bonding, (3) hydrogen bonding, (4) charge transfer, (5) ligand-exchange and ion bonding, (6) direct and induced ion-dipole and dipole-dipole interactions, and (7) chemisorption by the formation of a covalent bond. Sorption of organic pollutants sometimes can be explained with the simultaneous contribution of two of more of these mechanisms, especially when the nonpolar or polar character of the compounds is not well defined.

4.3 Vapor Pressure

The vapor pressure (or equilibrium vapor pressure) of an organic compound is the pressure exerted by a vapor in thermodynamic equilibrium with the condensed phase (liquid or solid) at a given temperature in a closed system (Boethling and Mackay, 2000; Mackay et al., 2006). The equilibrium vapor pressure is an indication of the evaporation rate of the liquid and relates to the tendency of particles to escape from the liquid (or a solid). A substance with a high vapor pressure at normal temperatures is often referred to as volatile. As the temperature of a liquid increases, the kinetic energy of its molecules also increases and the number of molecules passing from the liquid phase to the vapor phase also increases, thereby increasing the vapor pressure.

Vapor pressure controls the volatility of a chemical from soil, and along with its water solubility, determines evaporation from water (Boethling and Mackay, 2000). Predicting the volatility of a chemical in soil systems is important for estimating chemical partitioning in the subsurface between sorbed phase, dissolved phase, and gas phase. Vapor pressure is an important parameter in estimating vapor transport of an organic chemical in air. It is often useful to determine and other properties of a chemical by assuming that it is a liquid or supercooled liquid at a temperature less than the melting point (Mackay et al., 2006). At very low environmental concentrations such as in liquid solutions or on aerosol particles, pure chemical behavior relates to the liquid rather than the solid state.

The vapor pressure of any substance increases nonlinearly with temperature according to the Clausius-Clapeyron relation, which equation allows an expression of the pressure, P, the enthalpy of vaporization, DHvap, and the temperature, T, as:

P=AexpDHvap/RT

si19_e

In this equation, R (= 8.3145 J mol− 1 K− 1) and A are the gas constant and unknown constant, respectively. If P1 and P2 are the pressures at two temperatures T1 and T2, the equation has the form:

lnP1/P2=DHvap/R1/T11/T2

si20_e

The Clausius-Clapeyron equation allows an estimate of the vapor pressure at another temperature, if the vapor pressure is known at some temperature, and if the enthalpy of vaporization is known.

The atmospheric pressure boiling point (the normal boiling point) of a liquid is the temperature at which the vapor pressure equals the ambient atmospheric pressure. With any incremental increase in that temperature, the vapor pressure becomes sufficient to overcome atmospheric pressure and lift the liquid to form vapor bubbles inside the bulk of the substance. Bubble formation deeper in the liquid requires a higher pressure, and therefore higher temperature, because the fluid pressure increases above the atmospheric pressure as the depth increases.

4.4 Volatility

Volatility of an organic compound or of the organic components of a mixture is an important loss mechanism in the overall materials balance. The key environmental factors affecting volatilization are the reaction constant (surface transfer rate of dissolved oxygen per mixed depth of the water body), wind speed, and the mixed depth of the water. Furthermore, when evaluating the volatilization of complex mixtures vs. single chemicals, only physicochemical properties can be affected differently by mixtures. Thus, only by altering either aqueous solubility or vapor pressure of a chemical by interactions with other chemicals can volatilization rates be altered. Also, if the composition of a given mixture is known as well as the solubility of each constituent (from knowledge of the physiochemical properties of each constituent), it is possible to anticipate that rate of change for volatilization due to chemical interactions.

Organic chemicals, whether they are produced by the chemical industry or by the pharmaceutical industry or the petroleum industry, can be considered as environmentally transportable materials, the character of which is determined by several chemical and physical properties (i.e., solubility, vapor pressure, and propensity to bind with soil and organic particles). These properties are the basis of measures of leachability and volatility of individual hydrocarbons. Thus, the transport or organic chemicals either as individual chemicals or as mixtures (such as the various crude oil-derived products) can be considered by equivalent carbon number to be grouped into thirteen different fractions. The analytical fractions are then set to match these transport fractions, using specific n-alkane derivatives to mark the analytical results for aliphatic compounds and selected aromatic compounds to delineate hydrocarbons containing benzene rings.

Although organic chemicals grouped by transport fraction generally have similar toxicological properties, this is not always the case. For example, benzene is a carcinogen but many alkyl-substituted benzenes do not fall under this classification. However, it is more appropriate to group benzene with compounds that have similar environmental transport properties than to group it with other carcinogens such as benzo(a)pyrene that have very different environmental transport properties. Nevertheless, consultation of any reference work that lists the properties of chemicals will show the properties and hazardous nature of the types of chemicals that are found in petroleum. In addition, petroleum is used to make petroleum products, which can contaminate the environment.

The range of chemicals produced by the organic chemicals industry and by the petroleum refining industry is so vast that summarizing the properties and/or the toxicity or general hazard of crude oil in general or even for a specific crude oil is a difficult task. However, petroleum and some petroleum products, because of the hydrocarbon content, are at least theoretically biodegradable but large-scale spills can overwhelm the ability of the ecosystem to break the oil down. The toxicological implications from petroleum occur primarily from exposure to or biological metabolism of aromatic structures. These implications change as an oil spill ages or is weathered.

4.4.1 Low-Boiling Organic Chemicals

Many of the gaseous organic chemicals and liquid organic chemicals that are low-boiling materials (including crude oil and crude oil-delved products) fall into the class of chemicals which have one or more of the following characteristics that are considered to be hazardous by the Environmental Protection Agency: (1) ignitability, (2) flammability, (3) corrosivity, and (4) reactivity. In summary, many of the specific chemicals in crude oil and crude oil-derived products are hazardous because of their chemical reactivity, fire hazard, toxicity, and other properties. In fact, a simple definition of a hazardous chemical (or hazardous waste) is that it is a chemical substance (or chemical waste) that has been inadvertently released, discarded, abandoned, neglected, or designated as a waste material and has the potential to be detrimental to the environment. Alternatively, a hazardous chemical may be a chemical that may interact with other (chemical) substances to give a product that is hazardous to the environment. Low-boiling organic chemicals (whether crude oil-derived or derived from another source) fit very well into this definition; examples of some commonly encountered volatile hydrocarbons are

Aliphatic HydrocarbonsAromatic Hydrocarbons
Pentane derivativesBenzene
Hexane derivativesToluene
Heptane derivativesEthylbenzene
Octane derivativesXylene isomers and derivatives
Nonane derivativesNaphthalene derivatives
Decane derivativesPhenanthrene derivatives
Anthracene derivatives
Acenaphthylene derivatives

For example, a liquid that has a flash point of < 60°C (140°F) is considered ignitable. Some examples are: benzene, hexane, heptane, benzene, pentane, petroleum ether (low boiling), toluene, and the xylene isomers. An organic chemical is classed as flammable if the chemical has the ability of a substance to burn or ignite, causing fire or combustion. The degree of difficulty required to cause the combustion of a substance is quantified through standard test methods (Speight, 2001, 2014, 2015). The data from such test methods are used in regulations that govern the storage and handling of highly flammable substances inside and outside of structures and in surface and air transportation.

An aqueous solution that has a pH of ≤ 2, or ≥ 12.5 is considered corrosive. Most organic chemicals, crude oils, and crude oil-derived products are not corrosive but many of the chemicals used in refineries are corrosive—corrosive materials include inorganic chemicals such as sodium hydroxide as well as other acids or bases. The term reactivity applies to chemicals that react violently with air or water and, as a result, are considered to be hazardous chemicals. Reactive organic chemicals include chemicals capable of detonation (TNT) when subjected to an initiating source.

Gas condensate also falls within the VOC category and condensate release can be equated to the release of volatile constituents but are often named as such because of the specific constituents of the condensate, often with some reference to the gas condensate that is produced by certain crude oil wells and natural gas wells. However, the condensate is often restricted to the low-boiling alkane derivative as well as benzene, toluene, ethyl benzene, and the xylene isomers (BTEX).

4.4.2 High-Boiling Organic Chemicals

In almost all cases of hydrocarbon contamination, attention must be directed to the presence of semivolatile hydrocarbon derivatives and nonvolatile hydrocarbon derivatives.

Among the polynuclear aromatic hydrocarbons, the toxicity of many hydrocarbon liquids (especially crude oils and crude oil-derived products) is a function of its di- and tri-aromatic hydrocarbon content. Like the single aromatic ring variations, including benzene, toluene, and the xylene isomers, all are relatively volatile compounds with varying degrees of water solubility. However, in the higher boiling hydrocarbons liquids (particular products designed as fuel oil), the two-ring condensed aromatic hydrocarbons, naphthalene and the various homologs, are less acutely toxic than benzene but are more prevalent for a longer period after a spill or discharge. The toxicity of different crude oils and refined products (such as naphtha and naphtha-derived solvents and fuels) depends not only on the total concentration of hydrocarbons but also on the hydrocarbon composition in the water-soluble fraction (WSF) as well as on the (1) the degree of water solubility, (2) the concentrations of the individual component, and (3) the toxicity of the components either individually or collectively. The WSFs prepared from different crude oils will vary in terms of these three parameters. WSFs of the refined products (such as naphtha and naphtha-derived solvents, and fuels such as No. 2 fuel oil and Bunker C oil) are more toxic to the floral and faunal inhabitants of many ecosystems than the typical WSF of crude. Organic chemical either having a higher number of condensed rings or with methyl substituents on the rinds are typically more toxic than the less substituted derivatives but tend to be less water soluble and thus less plentiful in the WSF. There are also indications that pure naphthalene (a constituent of moth balls that are, by definition, toxic to moths) and alkyl naphthalene derivatives are from three-to-ten times more toxic to test animals than are benzene and alkylbenzene derivatives. In addition, and because of the low water-solubility of tricyclic and polycyclic (polynuclear) aromatic hydrocarbons (that is, those aromatic hydrocarbons heavier than naphthalene), these compounds are generally present at very low concentrations in the WSF of oil. Therefore, the results of this study and others conclude that the soluble aromatics of crude oil (such as benzene, toluene, ethylbenzene, xylene isomers, and naphthalene derivatives) produce the majority of toxic effects of crude oil in the environment.

The higher molecular weight aromatic structures (with four to five condensed aromatic rings), which are the more persistent in the environment, have the potential for chronic toxicological effects. Since these compounds are nonvolatile and are relatively insoluble in water, their main routes of exposure are through ingestion and epidermal contact. Some of the compounds in this classification are considered possible human carcinogens; these include benzo(a)pyrene, benzo(e)pyrene, benzo(a)anthracene, benzo(b, j, and k)fluorene, benzo(ghi)perylene, chrysene, dibenzo(ah)anthracene, and pyrene.

4.5 Water Solubility

The water solubility of an organic compound is the solubility of the organic compound in water (typically in grams per liter) at a designated temperature and pressure. At the molecular level, solubility is controlled by the energy balance of intermolecular forces between solute-solute, solvent-solvent, and solute-solvent molecules. Intermolecular forces vary in strength from very weak induced dipole to the much stronger dipole-dipole forces (such as hydrogen bonding). Most organic molecules are relatively nonpolar and are usually soluble in organic solvents (e.g., diethyl ether, dichloromethane, chloroform, organic chemicals ether, hexanes, etc.) but not in polar solvents such as water. However, some organic molecules are more polar and therefore soluble in water, is generally indicative of a high ratio of polar group(s) to the nonpolar hydrocarbon chain, i.e., a low molecular weight compound containing a hydroxyl group (glyph_sbndOH) or an amino group (glyph_sbndNH2) or a carboxylic acid group (glyph_sbndCO2H group) or a higher molecular weight compound containing several polar groups. The presence of an acidic carboxylic acid group or a basic amino group in a water-soluble compound can be detected by the low or high pH, respectively, of the solution.

Water solubility is one of the most important properties for evaluating the fate and direct measure of chemical since chemicals with high water solubility will partition readily and rapidly into the aqueous phase and will often remain in solution and be available for degradation. Chemicals that are sparingly soluble in water will often dissolve slowly into solution and partition more readily into other phases including air, solids, and the surface of solid particles including soil (Boethling and Mackay, 2000). Water solubility is used to determine the maximum theoretical concentration of a chemical in the soil pore water and is also important for estimating the air-water partitioning coefficient (Kaw), used to determine the partitioning behavior of an organic chemical in the soil.

There is a simple, very useful, and practical empirical rule that is quite reliable—like dissolves like—and it is based on the polarity of the systems insofar as polar molecules dissolve in polar solvents (such as water, alcohols) and nonpolar molecules dissolve in nonpolar solvents (such as liquid hydrocarbons).

The solubility of organic chemicals is an important property when assessing toxicity since the water solubility of an organic chemical determines the routes of exposure that are possible. In fact, an easy way of assessment of the water solubility of organic chemicals is that the solubility of an organic chemical is approximately inversely proportional to molecular weight—lower molecular weight hydrocarbon derivatives (excluding the presence of polar functional groups) are typically more soluble in water than the higher molecular weight compounds. Thus, lower molecular weight hydrocarbons (specifically, the C4 to C8 alkanes, including the aromatic compounds) are relatively soluble, up to approximately 2000 ppm, while the higher molecular weight hydrocarbon derivatives are much less soluble in water. Typically, the most soluble components are also the most toxic but whether this toxicity is due to the chemical and structural aspects of the lower molecular weight derivatives or whether it is noticed more frequently because of the relative ease of transportation is subject to debate.

Finally, each functional group has a particular set of chemical properties that allow it to be identified. Some of these properties can be demonstrated by observing solubility behavior, while others can be seen in chemical reactions that are accompanied by color changes, precipitate formation, or other visible effects. The identification and characterization of the organic chemicals is an important aspect of environmental organic chemistry. Although it is often possible to establish the structure of spectroscopic test methods (Chapter 5), the spectroscopic data should be supplemented with additional information such as (1) the physical state and (2) the relevant properties such as melting point, boiling point, solubility, odor, elemental analysis, and confirmatory tests for the presence of functional groups. The later test methods usually involve simple laboratory tests for solubility from which conclusions can be drawn which, when combined with other test data, can provide valuable information about the nature of the sample, but whether or not this can be applied to complex mixtures is another issue. Nevertheless, the solubility of an organic compound in water, dilute acid, or dilute base can provide useful information about the presence or absence of certain functional groups. For example, (1) solubility in water, (2) solubility in sodium hydroxide, and (3) solubility in hydrochloric acid (Fig. 6.1).

f06-01-9780128044926
Fig. 6.1 General schematic for testing the solubility of organic chemicals.

Most organic compounds are insoluble in water, except for low molecular weight amines and oxygen-containing compounds. Low molecular weight compounds are generally limited to those with fewer than five carbon atoms. Carboxylic acids (RCO2H) with fewer than five carbon atoms in the molecule are soluble in water and form solutions that give an acidic response (pH < 7) when tested with litmus paper. Amines (RNH2) with fewer than five carbon atoms in the molecule are also soluble in water, and their amine solution gives a basic response (pH > 7) when tested with litmus paper. Ketones, aldehydes, and alcohols with fewer than five carbon atoms are soluble in water and form neutral solutions (pH 7).

In addition, solubility in sodium hydroxide solution (usually 6 M NaOH, 6 molar NaOH) is a positive identification test for acids and acid derivatives. A carboxylic acid that is insoluble in pure water will be soluble in base due to the formation of the sodium salt of the acid as the acid is neutralized by the base. Also, solubility in hydrochloric acid (usually 6 M HCl, 6 molar HCl) is a positive identification test for bases. Amines that are insoluble in pure water will be soluble in hydrochloric acid due to the formation of an ammonium chloride-type salt (RNH3+ Cl).

4.6 Total Petroleum Hydrocarbons

Within the environmental arena, there is often reference to total petroleum hydrocarbons (TPHs) which is used because; given the wide variety of chemicals in petroleum and in petroleum products, it is not practical to consider each constituent separately. After an incident, it is more usual to measure the amount of TPHs at the site. The term TPHs is used environmentally to describe the family of several hundred chemical compounds that originally come from petroleum (Speight, 2001, 2014, 2015). Chemicals that may be included in the TPHs are hexane, heptane (and higher molecular weight homologs), benzene, toluene, xylene isomers (and the higher molecular weight homologs), and naphthalene as well as the constituents of other petroleum products such as gasoline and diesel fuel. It is likely that samples of the TPHs collected at a specific site will contain only some, or a mixture, of these chemicals but samples from different sites cannot be (should not be) expected to be the same in terms of composition and content of chemicals because of the variations in the composition of the starting crude oil feedstocks.

Petroleum products, themselves, are the source of many components, but do not adequately define TPHs. However, the composition of petroleum products assists in understanding the hydrocarbon derivatives that become environmental contaminants, but any ultimate exposure is determined also by how the product changes with use, by the nature of the release, and by the hydrocarbon's environmental fate. When petroleum products are released into the environment, changes occur that significantly affect their potential effects. Physical, chemical, and biological processes change the location and concentration of hydrocarbons at any particular site.

Hydrocarbons (in this context, petroleum-derived hydrocarbons) may enter the environment through accidents, from industrial releases, or as by-products from commercial or private uses such as direct release into water through spills or leaks. When release into water occurs, some of the hydrocarbons float on the water and form surface films while others may sink and form bottom sediments. Bacteria and microorganisms in the water have the potential to break down some of the hydrocarbons over varying periods of time that are dependent on the ambient conditions. On the other hand, hydrocarbon derivatives that are spilled on to the soil may remain for a long time, subject to the molecular size and volatility.

In addition, the amount of TPHs is the measurable amount of petroleum-based hydrocarbon in an environmental medium, whether it is air, water, or land. It is, thus, dependent on analysis of the medium in which it is found and since it is a measured, gross quantity without identification of its constituents, the TPHs’ data still represent a mixture. Thus, the data derived from measurement of the petroleum hydrocarbons in a particular environment is not a direct indicator of risk to humans or to the environment. The data may even be the results from one of several analytical methods, some of which have been used for decades and others developed in the past several years.

Analysis for TPHs (EPA Method 418.1, 2016) provides a one number value of the petroleum hydrocarbons in a given environmental medium. It does not, however, provide information on the composition (i.e., individual constituents) of the hydrocarbon mixture. The amount of hydrocarbon contaminants measured by this method depends on the ability of the solvent used to extract the hydrocarbon from the environmental media and the absorption of infrared light (infrared spectroscopy) by the hydrocarbons in the solvent extract. The method is not specific to hydrocarbon derivatives and does not always indicate petroleum contamination since humic acid, a nonpetroleum material and a constituent of many soils, can be detected by this method. Another analytical method commonly used for TPHs (EPA Method 8015 Modified, 2016) gives the concentration of purgeable and extractable hydrocarbons. These are sometimes referred to as gasoline range organics and diesel range organics because the boiling point ranges of the hydrocarbon in each roughly correspond to those of gasoline (C6 to C10–12) and diesel fuel (C8–12 to C24–26), respectively. Purgeable hydrocarbons are measured by purge-and-trap gas chromatography analysis using a flame ionization detector, while the extractable hydrocarbons are extracted and concentrated prior to analysis. The results are most frequently reported as single numbers for purgeable and extractable hydrocarbon derivatives. Another method (based on EPA Method 8015 Modified, 2016) gives a measure of the aromatic and aliphatic content of the hydrocarbon in each of several carbon number ranges (fractions). An important feature of the analytical methods for the TPHs is the use of an equivalent carbon number index. This index represents equivalent boiling points for hydrocarbons and is the physical characteristic that is the basis for separating petroleum (and other) components in chemical analysis and for identifying the ratios of specific hydrocarbon derivatives in the sample from which a source might be deduced.

References

Andrews J.E., Brimblecombe P., Jickells T.D., Liss P.S. An Introduction to Environmental Chemistry. Oxford: Blackwell Science Publications; 1996.

Bergen B.J., Nelson W.G., Pruell R.J. Bioaccumulation of PCB congeners by blue mussels (Mytilus edulis) deployed in New Bedford Harbor, Massachusetts. Environ. Toxicol. Chem. 1993;12:1671–1681.

Boethling R.S., Mackay D. Handbook of Property Estimation Methods. Boca Raton, FL: Lewis Publishers; 2000.

Calvert J., ed. The Chemistry of the Atmosphere: Its Impact on Global Change. Oxford: Blackwell Scientific Publications; 1994.

Cole J.G., Mackay D. Correlating environmental partitioning properties of organic compounds: the three solubility approach. Environ. Toxicol. Chem. 2000;19(2):265–270.

Corvalán C., Kjellström T. Health and environment analysis for decision-making. In: Briggs D., Corvalán C., Nurminen M., eds. Linkage Methods for Environment and Health Analysis—General Guidelines. Geneva: United Nations Environment Program (in cooperation with the World Health Organization and the United States Environmental Protection Agency); 1996:1–18.

Crompton T.R. Organic Compounds in Soils, Sediments and Sludge: Analysis and Determination. Boca Raton, FL: CRC Press, Taylor & Francis Group; 2012.

Dean J.R. Extraction Methods for Environmental Analysis. New York, NY: John Wiley & Sons, Inc.; 1998.

Delle Site A. Factors affecting sorption of organic compounds in natural sorbent/water systems and sorption coefficients for selected pollutants. A review. J. Phys. Chem. Ref. Data. 2001;30(1):187–439.

EPA Method 418.1. Petroleum Hydrocarbons: Total Recoverable. Washington, DC: United States Environmental Protection Agency; 2016.

EPA Method 8015 Modified. Purgeable and Extractable Hydrocarbons. Washington, DC: United States Environmental Protection Agency; 2016.

Firor J. The Changing Atmosphere. New Haven, CT: Yale University Press; 1990.

Goody R. Principles of Atmospheric Physics and Chemistry. Oxford: Oxford University Press; 1995.

Lyman W.J., Reehl W.F., Rosenblatt D.H., eds. Handbook of Chemical Property Estimation Methods. New York, NY: McGraw-Hill; 1982.

Lyman W.J., Reehl W.F., Rosenblatt D.H. Handbook of Chemical Property Estimation Methods. Washington, DC: American Chemical Society; 1990.

Mackay D. Correlation of Bioconcentration Factors. Environ. Sci. Technol. 1982;16:274–278.

Mackay D., Shiu W.-Y., Ma K.-C., Lee S. Handbook of Physical-Chemical Properties and Environmental Fate of Organic Chemicals. second ed. Boca Raton, FL: CRC Press, Taylor & Francis Group; 2006.

Manahan S.E. Environmental Chemistry. ninth ed. Boca Raton, FL: CRC Press, Taylor & Francis Group; 2010.

McCall P.J., Laskowski D.A., Swann R.L., Dishburger H.J. Estimation of environmental partitioning of organic chemicals in model ecosystems. In: Gunther F.A., Gunther J.D., eds. New York, NY: Springer; 231–244. Residue Reviews. 1983;vol. 85.

Patnaik P., ed. Dean's Analytical Chemistry Handbook. second ed. New York, NY: McGraw-Hill; 2004.

Schwarzenbach R.P., Gschwend P.M., Imboden D.M. Environmental Organic Chemistry. second ed. Hoboken, NJ: John Wiley & Sons Inc.; 2003.

Smith R.K. Handbook of Environmental Analysis. fourth ed. Schenectady, NY: Genium Publishing; 1999.

Speight J.G. Handbook of Petroleum Analysis. Hoboken, NJ: John Wiley & Sons Inc.; 2001.

Speight J.G. Environmental Analysis and Technology for the Refining Industry. Hoboken, NJ: John Wiley & Sons Inc.; 2005.

Speight J.G. The Chemistry and Technology of Petroleum. fifth ed. Boca Raton, FL: CRC Press, Taylor & Francis Group; 2014.

Speight J.G. Handbook of Petroleum Products Analysis. second ed. Hoboken, NJ: John Wiley & Sons Inc.; 2015.

Speight J.G., Islam M.R. Peak Energy—Myth or Reality. Salem, MA: Scrivener Publishing; 2016.

Spellman F.R. Handbook of Environmental Engineering. Boca Raton, FL: CRC Press, Taylor & Francis Group; 2016.

WMO. Scientific Assessment of Stratospheric Ozone 1991. WMO Global Ozone Research and Monitoring Project, Report No. 25 Geneva: World Meteorological Organization; 1992.

..................Content has been hidden....................

You can't read the all page of ebook, please click here login for view all page.
Reset