You are on page 1of 53

Chapter.VI.Environmental Hydraulics 1.

Definition Environmental hydraulics is the examination of the physical, chemical and biological attributes of flowing water, with the objective of protecting and enhancing the quality of the environment, including public welfare. (Canadian Hydraulics Centre) 2. Introduction to pollutants transport Processes which move pollutants and other compounds through the air, surface water, or subsurface environment or through engineered systems (for example, treatment reactors) are of particular interest to environmental engineers and scientists. Pollutant transport acts to move pollutants from the location at which they are generated, resulting in impacts which can be distant from the pollution source. On the other hand some pollutants, such as sewage sludge, can be degraded in the environment if they are sufficiently dilute. For these pollutants, slow transport---slow dilution---can result in excessively high pollutant concentrations, with resulting increased adverse impacts. In this chapter, we discuss the processes which distribute pollutants in the environment. The goals of this discussion are twofold: To provide an understanding of the processes which cause pollutant transport And to present and apply the mathematical formulas used to calculate pollutant fluxes.

3. Units of Concentration Before beginning the main topics of this chapter, it is necessary to cover units of concentration which will be used throughout the chapter. Pollutant concentration is the most important determinant in almost all aspects of pollutant fate and transport in the environment and in engineered systems. Concentration is also the driving force which controls the rate of chemical reactions and pollutant effects, such as toxicity, are often determined by concentration. Concentrations of pollutants and other chemicals are routinely expressed in a variety of units. The choice of units to use in a given situation depends on the pollutant, where it is located (e.g., air, water, or soil), and often on what the measurement will be used for. It is therefore necessary to become familiar with the units used and methods for converting between different sets of units.

Most of the ways concentration is represented fall into the categories shown in Table

a) Mass Concentration Units Concentration units based on pollutant mass include mass pollutant/total mass and mass pollutant/total volume. Examples of these are shown above. In these descriptions, mA is used to represent the mass of the pollutant, referred to as compound A. Mass/mass Units: ppm. Parts-per-million by mass (ppm, or ppmm) is defined as the number of units of mass of pollutant per million units of total mass. That is,

This definition is equivalent to the following general formula, which is used to calculate ppmm concentration from measurements of pollutant mass in a sample of total mass mtotal:

Note that the factor in explicitly here:

in equation 2 is really a conversion factor. It has implicit units as shown

Similar definitions are used for the units ppb , ppt , and % by mass. That is, 1 ppb( part per billion)= 1 g pollutant per billion ( ) g total, so that the number of ppb in a sample is equal to .
2

Percent by mass is analogously equal to the number of g pollutant per 100 g total. Be cautious about interpreting ppt values---they can refer to either parts per thousand or parts per trillion ( ). Example . Concentration in soil. A 1.0 kg sample of soil is analyzed for the pollutant trichloroethylene (TCE). The analysis indicates that the sample contains 5.0 mg of TCE. What is the TCE concentration in ppm ? Solution:

Mass/volume Units: mg/l.

In water, units of mg/L are common. Note from Example.2 that, in water ppmm is equivalent to mg/l. This is because the density of pure water is approximately 1000 g/l. Most aqueous solutions encountered in environmental engineering and science are dilute, meaning that dissolved material does not add significantly to the mass of the water, and the total density remains approximately 1000 g/l. Example. . Concentration in water. One liter of water is analyzed and found to contain 5.0 mg TCE. What is the TCE concentration in mg/l and ppm ? Solution: The concentration in units of mg/l is obtained simply by dividing the measured mass of TCE by the volume of water. No conversion is necessary, since the measured quantities are already in mass/volume units:

To convert to ppm , which is a mass/mass unit, it is necessary to convert the volume of water to mass of water, by dividing by the density of 1000 g/l:

For concentrations in the atmosphere, it is common to use units of mass/m3 air. For example, mg/m3. a g/m3are common. Example. . Concentration in gas. What is the carbon monoxide concentration (expressed in g/m3)of a 10 l gas mixture which contains1.0x10-6 mole of CO? Solution: In this case, we are presented with measured quantities which are in units of moles pollutant/total volume. To convert to mass of pollutant/total volume, we must convert moles of pollutant to mass of pollutant. This requires multiplying by the molecular weight. The next two lines of the solution are simply unit conversions, from grams to micrograms and from liters to cubic meters.

Note that the molecular weight of CO (28 g) is equal to 12 (molecular weight of C) plus 16 (molecular weight of O). b) Volume/volume and moles/moles Units. Units of volume fraction or mole fraction are frequently used for gas concentrations. The most common volume fraction units are ppmv (ppm by volume) which is defined as

Note that here, again, the factor (volume fraction).

is really a conversion factor, with units of 10-6ppmv

Also common is ppbv (parts per by volume). Volume/volume units have the advantage that they are unchanged as gases are compressed or expanded. (Note that atmospheric concentrations expressed as g/m3) decrease as the gas expands, since the pollutant mass remains constant but the volume increases.) The Ideal Gas Law. To convert gas concentration between mass/volume and volume/volume, we use the Ideal Gas Law, which states that

(Pressure) x(Volume taken up) = (No. of moles) x(R, gas constant) x(Temperature in Kelvin or Rankin). R, the universal gas constant, is a useful constant to know and may be expressed in many different sets of units. Some of the most useful are displayed in Table 1. The gas constant may be expressed in a number of different sets of units---always include all the units in each term and cancel them out to ensure that you are using the correct value of R in equation 5.

The ideal gas law states that the volume taken up by a given number of molecules of any gas is the same, no matter what the molecular weight or composition of the gas, as long as the pressure

and temperature are constant. Equation 5 can be rearranged to show that the volume taken up by n moles of gas is equal to

At standard conditions (P=1 atm, T=273 K), one mole of any pure gas will fill a volume of 22.4L. As an exercise, use the values of R\ given in Table 1 with equation 6 to derive this result. Example. . Gas concentration in volume or mole fraction. A gas mixture contains 0.001 mole of sulfur dioxide ( ) and 0.999 moles of air. What is the concentration, expressed in units of ppm ? Solution: The concentration in ppmv is calculated using the formula

To solve, we can convert the number of moles of to volume using the ideal gas law (V=NRT/P), also convert the total number of moles to volume, and then divide the two:

Substituting these volume terms in the equation for ppm , we obtain

Note that the terms (RT/P)in example 4 cancel out. This demonstrates a fact that can save you effort in calculating volume fraction or mole fraction concentrations--- for gases, volume ratios and mole ratios are equivalent. This is clear from the ideal gas law (equation 6), since at constant temperature and pressure the volume taken up by a gas is proportional to the number of moles. So, the following two equations are equivalent:

Example. . Gas concentration conversion. The concentration of SO2 in air is 100 ppb . What is this concentration in units of g/m3? (Temperature is 2 C and pressure is 1 atm. Remember that T ( K) is equal to T ( C) plus 273.) Solution: To accomplish this conversion, we will use the ideal gas law to convert volume of SO2 to moles of SO2, resulting in units of moles/l. These can be converted to g/m3using the molecular weight of SO2. First, we use the definition of ppb to obtain a volume ratio for :

We must now convert the volume of in the numerator to units of mass. This is done in two steps. First, we convert the volume to a number of moles, using the ideal gas law. From the standard form of the ideal gas law (PV=nRT), solving for the number of moles n\ yields . So, we multiply the quantity of given in volume units by P/RT to obtain units of moles. Note the choice of value for R used below, which is taken from Table 1, based on the units used in the problem.

For the second step, we convert the moles of and converting from g to g.

to mass of

, using the molecular weight of

Partial Pressures. Partial pressure units are also used to represent concentrations of gases. The total pressure exerted by a gas mixture may be considered as the sum of the partial pressures exerted by each component of the mixture. The partial pressure of each component is equal to the pressure that would be exerted if all of the other components of the mixture were suddenly removed. Partial pressure is commonly written as Pi, where i is the gas being considered. For example, the partial pressure of oxygen in the atmosphere at ground level may be written as atm.

The ideal gas law states that, at a given temperature and number of moles of gas, pressure is directly proportional to the number of moles of gas present:

Therefore, pressure fractions are identical to mole fractions (and volume fractions). For this reason, partial pressure may be calculated as the product of the mole or volume fraction and the total pressure. That is,

c) Moles/volume Units. Units of moles per liter (molarity, M) are often used to report concentrations of compounds dissolved in water. Molarity is defined as the number of moles of compound per liter of solution. Thus a M solution of nitric acid contains moles of per liter. Concentrations expressed in these units are read as molar. Thus, this solution would be described as `` \ molar.'' Example . . Molarity. Convert the concentration of TCE found in example 1.2 (5.0 ppm ) to units of M (molarity). (The molecular weight of TCE is 131.5 g.) Solution: In water, ppm is equivalent to mg/L, so we have 5.0 mg/l of TCE. Conversion to molarity units requires only the molecular weight:

Often, concentrations below 1 M are expressed in units of millimoles per liter, or millimolar (1 mM =1x10-3 moles/L). Thus, the concentration of TCE in example 1.6 is 0.038 mM. N.B:The units described here are the most common, but are not the only types of units you will encounter in environmental engineering problems. You might find additional ways to represent concentrations.

4. Mass and Energy Balances One of the fundamental laws of physics states that mass can neither be produced nor destroyed that is, mass is conserved. Equally fundamental is the law of conservation of energy. Although energy can change in form, it cannot be created or destroyed. These two laws of physics provide the basis for two tools which are used routinely in environmental engineering and science---the mass balance and the energy balance. This portion of the course deals with these tools. Mass balances are developed and applied in some detail in the following section, after which the concept of the energy balance is presented and applied.

Mass Balance

This principle of conservation of mass is extremely useful. It means that if the amount of a pollutant somewhere (say, in a lake) increases, then that increase cannot be the result of some ``magical'' formation out of nowhere. The pollutant must have been either carried into the lake from elsewhere or produced via chemical reaction from other compounds that were already in the lake. And, if chemical reactions produced the mass increase in our pollutant, they must also have caused a corresponding decrease in the mass of some other compounds. Thus, conservation of mass allows us to compile a budget of the mass of our pollutant in the lake. This budget keeps track of the amounts of pollutant entering the lake, leaving the lake, and the amount formed or destroyed by chemical reaction. This budget can be balanced for a given time period, similar to the way you might balance your checkbook:

Note that each term of this equation has units of mass. This form of balance is most useful when there is a clear beginning and end to the balance period, so that is meaningful. For example, in a checkbook balance is usually one month. In environmental problems, however, it is usually more convenient to work with values of mass flux---the rate at which mass enters or leaves a system. To develop an equation in terms of mass flux, the mass balance equation is divided by to produce an equation with units of mass per unit time. Dividing equation 12 by and moving the first term on the right (mass at time t) to the left hand side yields the following equation.

10

Note that each term in this equation has units of mass/time. The left hand side of equation 13 is equal to . In the limit as , this becomes dm/t, the rate of change of pollutant mass in the lake. We will refer to dm/dt as the accumulation rate of the pollutant. As , the first two terms on the right side of equation 13 become the mass flux into the lake and the mass flux out of the lake. The last term of equation 13 is the rate of chemical production or loss. To stress the fact that each term in the new equation refers to a flux or rate, we will use the symbol to refer to a mass flux with units of mass/time. The equation for mass balances is then

Equation 14 is the governing equation for the mass balances we will work with in this course. In the remainder of this section, we will examine the importance of carefully defining the region over which the mass balance is applied and discuss the terms of equation 14. We will then present examples of the main types of situations for which mass balances are useful.

The Control Volume

A mass balance is only meaningful in terms of a specific region of space, which has boundaries across which the terms and are determined. This region is called the control volume.
11

In our derivation of the mass balance equation, we have referred to the mass of pollutant in a lake and the fluxes of pollutant into and out of the lake---that is, we have used a lake as our control volume. Theoretically, any volume of any shape and location can be used. Realistically, however, certain control volumes are more useful than others. The most important attribute of a control volume is that it have boundaries over which you can calculate Terms of the Mass Balance Equation for a CSTR and .

A well-mixed tank is an analogue for many control volumes used in environmental engineering. For example, in our lake example it might be reasonable to assume that pollutants dumped into the lake are rapidly mixed throughout the entire lake. In environmental engineering and chemical engineering, the term Continuously Stirred Tank Reactor, or CSTR is used for such a system. An example of a CSTR is shown in Figure 1. We will use a mass balance for a control volume which encloses the CSTR in Figure 1 as an example to describe the meaning of each term in equation 14.

Mass Accumulation Rate,.

The mass accumulation rate is, by definition, , or . The total mass in the CSTR cannot usually be measured. For example, if the CSTR represented an entire lake, measuring the total pollutant mass would require analyzing all of the water in the lake. However, our assumption that the CSTR is well-mixed means that this is not necessary. If the tank is wellmixed, then the concentration of our pollutant is the same everywhere in the tank, and we need only to measure the concentration in a sampling from the tank. Using concentration units of (mass)/(volume), the total pollutant mass in the tank is equal to , where V is the volume of the CSTR. Thus, the accumulation rate is equal to

12

Here, we have made the assumption that the volume of the CSTR is constant. This is usually a reasonable assumption for liquids, although it may not always be valid for gases. However, will always be equal to d(VC)/dt. Mass balance problems can be divided into those that are in steady state and those that are nonsteady state. A steady-state situation is one in which things do not change with time---the incoming concentration and flow rate are constant, the outgoing flow rate is constant, and therefore the concentration in the control volume is constant. For steady-state systems, then, . Non-steady-state conditions result whenever flows start or stop, or when the concentration in an incoming stream changes. For non-steady-state situations, Mass Flux in,. is nonzero.

The example in Figure 1 includes one pipe entering the CSTR. We will again use concentration measured in mass/volume units to calculate the flux entering the CSTR through the pipe. Often, we know the volumetric flow rate, Q, of each input stream. For the example of Figure 1 the pipe has a flow rate of Qin, with corresponding pollutant concentration of Cin. The mass flux is then given by

If it is not immediately clear how one knows that QxC gives a mass flux, consider the units of each term:

If the volumetric flow rate is not known, it may be calculated from other parameters. For example, if the fluid velocity v and the cross-sectional area A of the pipe are known, then . Another way to describe the flux is in terms of a flux density J times the area through which the flux occurs. J has units of mass/(erea.time), and we will study it in more detail when we cover diffusion in section 3. This type of flux notation is most useful at interfaces where there is no fluid flow, such as the interface between the air and water and the surface of a lake.

13

Mass Flux out, . The flux out of the CSTR is similarly equal to the product of volumetric flow rate in the exit pipe times the concentration in the exit pipe. Since the CSTR is well-mixed, the concentration in the liquid leaving the CSTR is equal to the concentration inside the CSTR. It is conventional to refer to the concentration within the CSTR simply as C. Thus,

Net Rate of Chemical Reaction, . The term refers to the net rate of production of our pollutant from chemical reactions,

in units of mass/time. Thus, if other compounds react to form our pollutant, will be greater than zero; if our pollutant reacts to form some other compounds, resulting in a loss of the pollutant, will be negative. Production or loss of a compound by a chemical reaction is usually described in terms of concentration, not mass. So it is necessary to multiply the chemical rate of change of concentration by the volume of the CSTR to obtain units of mass/time:

There are a number of possibilities for the form of The most common include:

, and the resulting

1. Conservative pollutant. Pollutants with no chemical formation or loss are called conservative because their mass is conserved without any corrections for chemistry. For such compounds, , implying that also. 2. O -order decay. The rate of loss of the pollutant is constant. For a pollutant with

order decay, and . 3. 1 -order decay. The rate of loss of the pollutant is directly proportional to its concentration: . For such a pollutant, . 4. Production at a rate dependent on the concentrations of other compounds in the CSTR. In this situation, our pollutant is produced by chemical reactions involving other compounds in the CSTR, and is greater than zero. Examples of this type of situation will be given in Part III of the course.

14

Reactor Analysis---the CSTR Reactor Analysis refers to the use of mass balances to analyze pollutant concentrations in a control volume which is a chemical reactor. Do not let the term ``reactor'' fool you, however. The reactor can be any control volume we want it to be. So the term reactor analysis is used to describe the application of the mass balance process to environmental situations also. Reactor analyses can be divided into two types: CSTRs (Continuously Stirred Tank Reactors) and PFRs, or Plug Flow Reactors. We have defined CSTRs already---they are simply well-mixed tanks which are used to model well-mixed environmental reservoirs. Plug Flow Reactors are essentially pipes, and they are used to model things like rivers, in which fluid is not mixed in the upstream-downstream direction. In this section, we will present examples of the types of situations CSTRs are used to model. Plug Flow Reactors are described and used in examples in the following section. Example 2.1 demonstrates the use of CSTR analysis to determine the concentration of a substance resulting from the mixing of two or more influent flows. This type of calculation will be used again in the third part of this course to determine the initial BOD loading in a river downstream of a sewage outflow. Examples 2.2 through 2.4 refer to the tank in Figure 1 and demonstrate steady-state and non-steady-state situations with and without first-order chemical decay. Calculations completely analogous to those in examples 2.2, 2.3, and 2.4 can be used to determine the concentration of sewage pollutants exiting a treatment reactor, the rate of increase of pollutant concentrations within a lake resulting from a new pollutant source, and the period required for pollutant levels to decay from a lake or reactor once the source is removed.

Example . . Mixing Problem

15

A sewage pipe from a wastewater treatment plant discharges 1.0 m3/s of effluent containing 5.0 mg/L of phosphorus compounds (reported as mg P/L) into a river with an upstream flow rate of 25 m /s and a phosphorus concentration of 0.010 mg P/l (see Figure 2). What is the resulting concentration of phosphorus in the river downstream of the sewage outflow, in units of mg/L?

Solution: To solve this problem, we will apply two mass balances to determine first the downstream volumetric flow rate (Qd/s) and, second, the downstream phosphorus concentration (Cd/s). But first, we must select a control volume. To ensure that the input and output fluxes cross the control volume boundaries, the control volume must cross the river upstream and downstream of the sewage outlet and must cross the sewage pipe. The selected control volume is shown in Figure 2 as a dotted line. We must assume that the control volume extends down the river far enough that the sewage and the river water become well-mixed before leaving the control volume. As long as that assumption is met, it makes absolutely no difference to our analysis how far downstream the control volume extends. Before beginning our analysis, we should determine whether this is a steady-state or non-steadystate problem, and whether the chemical reaction term will be nonzero. Since the problem statement does not refer to time at all, and it seems reasonable to assume that both the river and sewage have been flowing for some time and will continue to flow; this is a steady-state problem. Sewage does participate in chemical and biological reactions. However, we are interested here in mixing---that is, in what concentration results right after the two flows mix. So we will assume that the mixing occurs instantly, without sufficient time for any reactions to occur. (a) What is ? We will conduct a mass balance on the total river-water mass. In this case, the ``concentration'' of river water in mass/volume units is simply the density of the water, .

where the term has been set to zero because we are ignoring chemical reaction. Since this is a steady-state problem, dm/dt=0. Therefore, as long as the density is constant, (b) What is , or .

? Again, we have steady-state with no chemical formation or decay:

16

Plugging in, we find that

Example . . Steady-state CSTR with 1st-order Decay The CSTR shown in Figure 1 is used to treat an industrial waste, using a reaction which destroys the waste according to first-order kinetics: , where . The reactor volume is 500 m , the volumetric flow rate of the single inlet and exit is 50 m /day, and the inlet waste concentration is 100 mg/l. What is the outlet concentration? Solution: An obvious control volume is the tank itself. A single, constant outlet concentration is asked for and all problem conditions are constant. Therefore, this is a steady-state problem. The mass balance equation is

Solving for C, we find that

The numerical solution is


17

Example . . Non-steady-state CSTR, Conservative Substance The CSTR shown in Figure 1 is used with a conservative substance. The reactor is filled with clean water before it is started. After starting, waste containing a 100 mg/l of a pollutant is added at a flow rate of 50 m /day. The volume of the reactor is 500 m . What is the concentration exiting the reactor as a function of time after it is started?

Solution: The substance is conservative---therefore, there the chemical reaction term in the mass balance equation is equal to zero. The mass balance equation is

Because of the extra term on the right ( ), this equation cannot be immediately solved in the way that example 2.4 was solved. However, if we make a change of variables, we can make the form of this equation similar to that of example 2.4. Let . Since is constant,

. Therefore, the last equation above is equivalent to

Rearranging and integrating,

which yields

18

If we now substitute

for y, we obtain

The second equation is obtained using the observation that clean. Rearranging, we can obtain

, since the tank is started

This is the solution to the question posed in the problem statement. Note what happens as : \ and . This is not surprising, since this is a conservative substance. If we run the reactor for a long enough period, the concentration in the reactor will eventually reach the inlet concentration. Using the equation we have derived for C as a function of time, we could determine how long it would take for the concentration to reach, say, 90% of the inlet value. Example . . Non-steady-state CSTR with 1 -order Decay The manufacturing process that generates the waste in example 2.2 has to be shut down, and, starting at t=0, the concentration entering the CSTR is set to 0. What is the outlet concentration as a function of time after the concentration is set to 0? How long does it take the tank concentration to reach 10% of its initial, steady-state value?

Solution: This is clearly a non-steady-state problem, because conditions change as a function of time. To solve it, we will again use the tank as our control volume. The mass balance equation is

19

(a) To determine C as a function of time, we must solve the differential equation. Rearranging and integrating:

which yields

Since

is equal to

and, exponentiating both sides,

Plugging in the values from the problem, with 32 mg/l yields

equal to the steady-state solution of

(b)

20

How long will it take the concentration to reach 10% of its initial, steady-state value? That is, at what value of t is , we have ? At the time when

Taking the natural logarithm of both sides,

The Plug Flow Reactor

The Plug Flow Reactor (PFR) is used to model the chemical transformation of compounds as they are transported in ``pipes.'' The ``pipe'' may represent a river, a region between two mountain ranges through which air flows, or a variety of other conduits through which liquids or gases flow. Of course, it can even represent a pipe. A schematic diagram of a PFR is shown in Figure 3. As fluid flows down the PFR, the fluid is mixed in the radial direction, but mixing does not occur in the axial direction---each plug of fluid is considered a separate entity as it flows down the pipe. However, as the plug of fluid flows downstream, time passes. Therefore, there is an implicit time dependence even in steady-state PFR problems. However, because the velocity of the fluid in the PFR is constant, time and downstream distance are interchangeable: . We will use this observation together with the mass balance formulations we have worked with already to determine how pollutant concentrations vary during flow down a PFR.
21

To develop the equations which describe pollutant concentration in the plug of fluid as it flows down the PFR, we will conduct a mass balance on a control volume which encloses a section of the PFR of infinitesimally small thickness dx, as shown in Figure 4. Since the thickness is small, we can assume that the fluid in that region of the PFR is well-mixed. The mass balance equation for this control volume is

We have set equal to zero, indicating that this is a steady-state problem. We are assuming here that conditions at a given location in the PFR are constant. Concentrations can still vary along the PFR, however. Noting that the volume of our control volume is given by rearranging, we obtain , dividing by dx, and

In the limit as

, the left hand side becomes the derivative

, so we obtain

22

As discussed earlier, can take a variety of forms, depending on the type(s) of chemical reaction that are occurring.

No reaction. . From equation 20, this means that ---there is no variation of concentration along the pipe. (Of course, this result is obvious, since if there is no reaction the fluid is just moving along the pipe without changing in any way.) First-order reaction. obtain . Plugging into equation 20 for this case, we

which can be integrated as follows:

Therefore, for a PFR of length l,

where the volume of the PFR, V, is equal to the length times time area. Equation 26 describes the way in which concentration decreases during passage down a PFR with loss via a first-order reaction. Note that, since the time which passes during transport down the PFR is equal to , equation 26 is equivalent to

which is the solution to the differential equation which describes the loss of a pollutant by firstorder kinetics: . That is, in a plug flow reactor time and distance are interchangeable, and the concentration at any location in the PFR may be calculated simply by determining the chemical decay during the time it took to reach that location.
23

Comparison of the PFR to the CSTR.

The CSTR and the PFR are fundamentally different. When a parcel of fluid enters the CSTR, it is immediately mixed throughout the entire volume of the CSTR. In contrast, each parcel of fluid entering the PFR remains separate during its passage through the reactor. This difference results in differing behavior. We will look at these differences for one special case: the continuous addition of a pollutant to each reactor, with destruction of the pollutant within the reactor according to first-order kinetics. The two reactors are shown in Figure 5. We will assume that the incoming concentration ( ), the flow rate (Q), and the first-order reaction rate constant (k) are given and are the same for both reactors. Then, we will consider two common problems: (1) if we know the volume V (same for both reactors), what is the resulting outlet concentration ( )? and (2) if we need a specified outlet concentration, what volume of reactor is required? Table 2 summarizes the results of this comparison.

24

The results shown in Table 2 indicate that, for equal reactor volumes, the plug flow reactor is more efficient that the CSTR and, for equal outlet concentrations, a smaller PFR is required. Why is this? The answer has to do with the fundamental difference between the two reactors. In a PFR, each and every molecule spends the same amount of time in the reactor; that period is equal to . Since first-order decay occurs according to , the concentration in each parcel of fluid entering the reactor drops by this amount. In contrast, in a CSTR there is no single amount of time that each small parcel of fluid spends in the reactor. Some parcels may spend a long time mixing around inside the CSTR; other parcels may, by chance, reach the exit in a relatively short time. Since all these parcels are mixed together and result in a single outlet concentration, an average value of \ results.

To see why that average value is higher than the corresponding value for a PFR, consider what happens when is equal to 2, approximately the value in the first example of Table 2.
25

Then, . This is the value of \ that would result in the PFR. Let's assume that we can model the mixing in the CSTR by splitting the fluid entering the CSTR into two parcels. The first parcel remains in the CSTR only one quarter of the time a parcel would take to pass through the PFR, while the second parcel remains in the CSTR four times as long as it would in the PFR. (So the average time spent in the CSTR by the two parcels is the same as the time spent in the PFR---both are equal to .) The concentration in the first parcel when it reaches the CSTR exit is determined by its value of kt, which is 4 times larger than the value for the PFR: . The concentration in the second parcel is reduced less, because it spends a shorter time in the reactor: . The actual concentration in the exit of the CSTR in this situation would be the average of the concentrations in the two parcels, so .

Thus, the resulting value of for the CSTR is higher than that for the PFR (0.30 versus 0.14), even though the average residence time is the same for both reactors. The reason for this is illustrated in Figure 6, and results from the fact that concentration decays exponentially with time for a first-order reaction. Thus, the parcel that spends a shorter period of time in the CSTR exits with a concentration that is increased significantly relative to the PFR. However, the parcel that spends a longer period in the CSTR exits with a concentration that is decreased only a small amount (again, relative to the PFR).

26

Example . . Required Volume in a PFR Determine the volume required for a PFR to obtain the same degree of pollutant reduction as the CSTR in example 2.2. Assume that the flow rate and first-order decay rate constant are unchanged( , ).

Solution:

The

CSTR

in

example

2.2

achieved

pollutant

decrease

of

. From equation 26,

27

Solving for V, we obtain

As expected, this volume is smaller than the 500 m required for the CSTR in example 2.2.

Retention Time and Other Expressions for . A number of terms are used to describe the average period spent in a given reactor. The terms retention time, detention time, and residence time are all used to refer to , the average period spent in the reactor. This parameter has units of time. As discussed above, for a plug flow reactor the retention time is actually the time spent in the reactor. However, for a CSTR the retention time is the average period spent in the reactor. The reciprocal of the retention time, , has units of inverse time---the same units as a firstorder rate constant. This value is sometimes referred to as the exchange rate.

Example . . Retention Time in CSTR and PFR Calculate the retention times in the CSTR of example 2.2 and the PFR of example 2.5. Solution: For the CSTR in example 2.2,

For the PFR in example 2.5,

28

Example . . Retention Times for the Great Lakes Calculate the retention times for Lake Michigan and Lake Ontario using the data given in Table 3. Solution: Again, we can calculate the retention times as . Note that we are assuming that the lakes can be modeled as CSTRs (or PFRs, but a CSTR makes more sense for a lake). This assumption is not far off for retention times significantly longer than one year.

For

Lake

Michigan,

For Lake Ontaria, The higher flow and smaller volume of Lake Ontario results in a significantly shorter retention time. This means that pollutant concentrations can increase in Lake Ontario much more quickly than they can in Lake Michigan, but it also means that concentrations will drop much more quickly in Lake Ontario if a pollutant source is eliminated.

29

Energy Balance

Modern society is dependent on the use of energy. Such use requires transformations in the form of energy and control of energy flows. For example, when coal is burned at a power plant, the chemical energy present in the coal is converted to heat, which is then converted in the plant's generators to electrical energy. Eventually, the electrical energy is converted back into heat for warmth or used to turn motors. However, energy flows and transformation are also the cause of environmental problems. Thermal heat energy from electrical power plants can result in increased temperature in rivers used for cooling water; ``greenhouse'' pollutants in the atmosphere alter the energy balance of the earth and may cause significant increases in global temperatures in the future; and many of our uses of energy are themselves associated with emissions of pollutants. We can keep track of the movement of energy and changes in its form using energy balances, which are analogous to the mass balances we discussed in the previous section. We can do this because of the law of conservation of energy which states that energy can neither be produced nor destroyed. (Conservation of energy is sometimes referred to as the first law of thermodynamics.) As long as we consider all the possible forms of energy, there is no term in energy balances which is analogous to the chemical reaction term in mass balances. That is, we can treat energy as a conservative substance. Forms of Energy

The forms of energy can be divided two types: internal energy and external energy. Energy which is a part of the molecular structure or organization of a given substance is internal. Energy
30

which results from the location or motion of the substance is external. Examples of external energy include gravitational potential energy and kinetic energy. Gravitational potential energy is the energy gained when a mass is moved to a higher location above the earth. Kinetic energy is the energy which results from the movement of objects. When a rock thrown off of a cliff accelerates toward the ground, the sum of kinetic and potential energy is conserved (neglecting friction)---as it falls it loses potential energy, but increases in speed, gaining kinetic energy. Examples of some common forms of energy are given in Table 4. Heat is a form of internal energy. It results from the random motions of atoms. Heat is thus really a form of kinetic energy, although it is considered separately. When you heat a pot of water, you are adding energy to the water. That energy is stored in the form of internal energy, and the change in internal energy of the water is given by

where C is the heat capacity or specific heat of the water, with units of [energy]/([mass][temperature]). Heat capacity is a property of a given material. For water, the heat capacity is 1 BTU/( ), or 4184 J/( ).

Chemical internal energy reflects the energy in the chemical bonds of a substance. This form of energy is composed of two parts: 1. The strengths of the atomic bonds in the substance. When chemical reactions occur, if the sum of the energies of the products is less than that for the reactants, a reduction in chemical internal energy has occurred. As a result of the conservation of energy, this leftover energy must show up in a different form. Usually, the energy is released as heat. This fact is used to our advantage when we burn fuel. 2. The energy in the bonds between molecules. This energy depends on the phase of the material---whether it is a solid, liquid, or gas. The energy required to change phases is known as the latent heat. Values of the latent heat of fusion (the energy released when a material changes from liquid to solid phase) and the latent heat of condensation or vaporization (the energy released when a substance changes from gas to liquid phase) are tabulated for many substances. When a substance changes from solid to liquid or from liquid to gas, it gains internal energy. This energy must come from somewhere. For example, when water evaporates it takes up energy from its surroundings, and this is why evaporation of sweat cools us. Conducting an Energy Balance In analogy with the mass balance equation (equation 14), we will use the following equation to conduct energy balances:
31

We will illustrate the use of this relationship with some examples.

Example . . Heating water A 40-gallon electric water heater is used to heat tap water (temperature 50 F (10 C)). The heating level is set to the maximum level while several people take consecutive showers. If, at the maximum heating level, the heater uses 5 kW of electricity, and the water use rate is a continuous 2 gallons/min, what will be the temperature of the water exiting the heater? Assume that the system is at steady-state and that the heater is 100% efficient; that is, it is perfectly insulated and all of the energy used goes to heat the water.

Solution: Our control volume for this problem is the water heater. We note that (since the system is at steady-state), the internal energy of the water in the water heater is constant. The energy added is used to heat water entering the water heater to the temperature at the outlet. The energy flux into the water heater comes from two sources: the heat content of the water entering the heater and the electrical heating element. The energy flux out of the water heater is just the internal energy of the water leaving the system. There is no net conversion of other forms of energy. Therefore, equation 30 may be rewritten as

Each term of this equation is an energy flux, and has the units of (energy/time). To solve, we need to use the same units in each term. We will use the definition of watts: watts are defined as Joules/s. In addition, we need to convert the water flow rate (gallons/min) to mass of water per unit time, using the density of water. Combining the first and third terms we obtain

32

which is a cold shower! (You may have foreseen this answer if you have ever taken a shower after the hot water in the tank was used up by previous showerers.)

Example . . Heating water From the previous example, we see that if one wants a hot shower, it is necessary to wait until the water in the tank can be reheated. How long would it take the temperature to reach 130 F (54 C) if no hot water were used during the heating period and the water temperature started at 20 C? Solution: In this case, the only energy input is the electrical heat, and there is no energy leaving the tank. Therefore, the rate of increase in internal energy is equal to the rate that electrical energy is used:

We will solve this for

, given that

is equal to

C.

The two previous examples related to the controlled conversion and transfer of energy for a beneficial use. However, the use of energy for heat always results in some loss to the environment due to imperfect insulation, resulting in higher energy use or less heating than one would calculate. In addition, the second law of thermodynamics states that it is impossible to convert heat energy to work with 100% efficiency. Conversion of heat to work is essentially what is done in the generator of an electric power plant, and as a result a significant fraction of the energy released from fuel combustion is lost during the conversion. Modern large power
33

plants convert fuel energy to electricity with an overall efficiency in the 30--35% range. The next example looks at what happens to the heat energy that is not converted to electricity. Finally, example 2.11 considers the implications of another aspect of the burning of fossil fuels in power plants, vehicles, and for heating. Whenever fossil fuels are burned, carbon atoms in the fuel are converted to carbon dioxide ( ) and released into the atmosphere. As a result of this process, the concentration in the atmosphere is increasing at a rate of about 1 ppmv/year. Carbon dioxide contributes to the greenhouse effect, which is considered in example 2.11. Example . . Thermal Pollution.

A 1000 MW ( W) power plant is located next to a river and uses cooling water from the river to remove its waste heat. What is the resulting increase in river temperature? (The power plant has an overall efficiency of 33%. Assuming that all of the waste heat from the power plant is removed with cooling water and added to the adjacent river. The river flow rate is 100 m /s. Solution: This problem is very similar to example 2.8, in that we are adding a specified amount of heat to a flow of water, and need to determine the resulting temperature rise. To solve, we first need to determine the amount of energy added. The power plant produces 1000 MW of electricity, but is only 33% efficient, meaning that it uses 3000 MW of fuel energy ( ). The heat energy added to the river is the amount that is not converted to electricity, or (3000-1000=2000 MW). We can now write our energy balance over the region of the river to which the heat is added. We will use of the water upstream, and \ to represent the temperature

to represent the temperature after heating:

Rearranging, we obtain

The remainder of this problem is basically a problem of unit conversions. To obtain requires multiplication of the given river volumetric flow rate by the density of water (1000 kg/m ). We also use the heat capacity of water,
34

C. Thus,

Example .

. Earth's Energy Balance and the Greenhouse Effect.

The global average surface temperature of the earth is determined by a balance between the energy added to the earth by the sun and the energy radiated away by the earth to space. Greenhouse gases, both natural and anthropogenic (or, human-affected), affect this energy balance. In this example, we will calculate the global average temperature without greenhouse gases and show the effect which greenhouse gases have on the earth's energy balance.

Solution: We will write an energy balance, with our control volume as the entire earth. For this system, our goal is to calculate annual-average temperatures. Over time periods of 1 yr, it is reasonable to assume that the system is in steady state, so our energy balance is simply

The energy flux in is equal to the solar energy intercepted by the earth. At the earth's distance from the sun, the sun radiates 342 W/m . We will refer to this value as S. The earth intercepts an amount of energy equal to S times the cross-sectional area of the earth: because the earth reflects part of this energy back to space, value: . However,

is equal to only 70% of this

The second term, , is equal to the energy radiated to space by the earth. The energy emitted per unit surface area of the earth is given by Boltzmann's Law:

where

is Boltzmann's constant, equal to

. To obtain

, we

multiply this value by the total surface area of the earth, . (We use the total surface area of the sphere here because energy is radiated away from the earth during both day and night.)
35

We can now solve our energy balance by setting

equal to

Plugging in the values for S and T=255 K, or -18 C.

and taking the fourth root yields an average temperature of

This is too cold! In fact, the globally averaged temperature at the surface of the earth is much warmer: 287 K. The reason for the difference is the presence of gases in the atmosphere that absorb the infrared radiation emitted by the earth and prevent it from reaching space. We neglected these gases in our energy balance. However, if we denote the energy flux absorbed and retained by these gases by , we can then correct our value for :

The reduction in which results from greenhouse gas absorption is sufficient to cause the higher observed surface temperature. Clearly, this is largely a natural phenomenon---surface temperatures were well above 255 K long before people began burning fossil fuels. The main natural greenhouse gas is water vapor. However, increasing atmospheric concentrations of carbon dioxide and other gases emitted by human activities are increasing the value of . So far, this increase is approximately 2 W/m \ averaged over the entire earth, and projections indicate that the increase could be as high as 5 W/m over the next 50 years. According to our energy balance, this increase in is expected to result in an increase in the globally averaged temperature. (There is considerable uncertainty in the precise value of the resulting increase, however, due to a number of complexities that we have not considered.)

36

2.3. Mass Transport Processes Advection and Diffusion

Transport processes in the environment may be divided into two categories: advection and diffusion. Advection refers to transport with the mean fluid flow. For example, if the wind is blowing toward the east, advection will carry any pollutants present in the atmosphere toward the east. Similarly, if a bag of dye is emptied into the center of a river, advection will carry the resulting spot of dye downstream. In contrast, diffusion refers to the transport of compounds through the action of random motions. Diffusion works to eliminate sharp discontinuities in concentration and results in smoother, flatter concentration profiles. Advective and diffusive processes can usually be considered independently. In the example of a spot of dye in a river, while advection moves the center of mass of the dye downstream, diffusion spreads out the concentrated spot of dye to a larger, less concentrated region. Definition of the Mass Flux Density

In our calculations of advective and diffusive fluxes, we will calculate the flux density across an imaginary plane oriented perpendicular to the direction of mass transfer. The resulting mass flux density is defined as the rate of mass transferred across the plane per unit time per unit area, with units of (mass)/(time).(length)2. We will use the symbol J to represent the flux density. J represents the mass flux density, expressed as the rate per unit area at which mass is transported across an imaginary plane. J has units of [M]/[L2T]. The total mass flux across a boundary ( ) can be calculated from the flux density simply by multiplying J by the area of the boundary:

In the following sections, we will consider the flux density which results from advection and from diffusion. The symbol J will be used to represent the flux density in each case, whether the flux is a result of advection, diffusion, or a combination of both processes. Calculation of the Advective Flux The advective flux refers to the movement of a compound along with flowing air or water. The advective flux density depends simply on concentration and flow velocity.
37

The fluid velocity, v, is a vector quantity---it has both magnitude and direction, and the flux J refers to the movement of pollutant mass in the same direction as the fluid flow. We will generally define our coordinate system so that the x-axis is oriented in the direction of fluid flow. In this case, the flux J will reflect a flux in the x-direction, and we will generally ignore the fact that J is really a vector. Example1. Calculation of the advective flux density. A sewage pipe from a wastewater treatment plant discharges 1.0 m3/s of effluent containing 5.0 mg/L of phosphorus compounds (reported as mg P/L) into a river with an upstream flow rate of 25 m3/s and a phosphorus concentration of 0.010 mg P/L. Calculate the average flux density J of phosphorus downstream of the sewage pipe .The cross-sectional area of the river is 30 m2. The downstream phosphorus concentration (Cd/s)is 0.20mg/L. Solution: The average river velocity is . Using the definition of flux density (equation 32), we find:

Diffusion Diffusion results from random motions of two types: the random motion of molecules in a fluid, and the random eddies which arise in turbulent flow. Diffusion from the random molecular motion is termed molecular diffusion; diffusion which result from turbulent eddies is called turbulent diffusion or eddy diffusion. We will compare these two types of diffusion below. First, however, we consider why diffusion occurs. Fick's Law In this section, we will derive Fick's Law, the equation which is used to calculate the diffusive flux density, by analyzing the results of random motion of a hypothetical box of molecules. The purpose of this derivation is to provide a qualitative and intuitive understanding of the reason that diffusion occurs, and the derivation itself is useful only for that purpose. In problems where

38

it is necessary to actually calculate the diffusive flux, we will normally start at the end of this derivation---that is with the Fick's Law equation (equation 40). Consider a box which is initially divided into two parts, as shown in Figure below. Each side of the box has a height and depth of 1 unit, and a width of length . Initially, the left portion of the box is filled with 10 molecules of gas x and the right side is filled with 20 molecules of gas y, as shown in the top half of Figure. We now remove the divider, and observe what happens.

Fig. Diffusion example As we know, molecules are never stationary. All of the molecules in our box are constantly moving around, and at any moment they have some probability of crossing the imaginary line at the center of the box. We will check the box and count the molecules on each side every seconds; we will call the probability that a molecule crosses the central line during the period between observations k. Let's say for now that . So the first time we check the box, after a period , 20% of the molecules that were originally on the left will have moved to the right, and 20% of the molecules that were originally on the right will have moved to the left. When we count the molecules on each side, then, we will find the situation shown in the bottom of Figure 8, with 8 ``x'' molecules remaining on the left and 2 on the right, and 16 ``y'' molecules remaining on the right, 4 having moved to the left. If we assume that the mass of each molecule is 1 unit, we can calculate concentrations in units of mass/volume, and find that the concentration differences
39

between the two boxes have been reduced from 10 to 6 for molecule ``x'', and from 20 to 16 for molecule ``y.'' This is a fundamental result of diffusion---that concentration differences are reduced. We also observe that the movement of molecule ``x'' is essentially independent from that of ``y'', so the diffusion of each molecule can be considered as a separate problem. That is, we don't have to worry about molecule ``x'' when we are calculating the diffusion of molecule ``y'', or vice versa. So diffusion moves mass from regions of higher concentration to regions of lower concentration, and if left to continue indefinitely, it would result in equal concentrations on both sides of the box. We now need to find the flux density J which diffusion causes. For this calculation, we will again use the situation shown in Figure 8, with a probability of any molecule crossing the central boundary during a period \ equal to k. Since each molecule can be considered independently, we will analyze the movement of a single molecule type, say molecule ``y.'' Let ML be the total mass of molecule ``y'' in the left half of the box, and MR equal the mass in the right half. Since our box has unit height and depth, the area perpendicular to the direction of diffusion is one square unit. Thus, the flux density---the flux per unit area---is just equal to the rate of mass transfer across the boundary. The amount of mass transferred from the left to right in a single time step is equal to kML, while the amount transferred from right to left during the same period is kMR. Thus, the rate of mass flux from left to right across the boundary is equal to (kML - is kMR ) divided by , or

Since it is more convenient to work with concentrations than with total mass values, we need to convert this equation to concentration units. The concentration in each half of the box is given by

Thus, substituting

for the mass in each half of the box, the flux density is equal to

40

Finally, we note that as

, and therefore

(Note that the negative sign in this equation is simply a result of the convention that flux is positive when it flows from left to right, but the derivative is positive when concentration increases toward the right.) This equation states that the flux of mass across an imaginary boundary is proportional to the concentration gradient at the boundary. Since the resulting flux cannot depend on arbitrary values of or , the product k(x)2/t must be constant. This product is the value we call the Diffusion Coefficient, D. Thus, we obtain Fick's Law:

The units of the diffusion coefficient are clear from an analysis of the units of equation 40 or from the units of the parameters in equation 39; the diffusion coefficient has the same units as . Since k is a probability, and thus has no units, the units of D must be (length)2 /(time). Diffusion coefficients are commonly reported in cm2/s. Before we go on, note the form of equation 40:

This form of equation will also appear later when we discuss Darcy's Law, which governs the rate at which groundwater flows through soil pores. The same equation also governs heat transfer, if we replace the concentration gradient with a temperature gradient. Molecular Diffusion

41

The molecules-in-a-box analysis used above is essentially an analysis of molecular diffusion. Purely molecular diffusion is relatively slow. Typical values of the diffusion coefficient D are approximately 10-2-10-1cm2/s for gases, and much lower, around cm /s for liquids. The difference in diffusion coefficient between gases and liquids is understandable if we consider that gas molecules are free to move much greater distances before being stopped by ``bumping'' into another molecule. The diffusion coefficient also varies with temperature and the molecular weight of the diffusing molecule. This is because the average speed of the random molecular motions is dependent on the kinetic energy of the molecules. As heat is added to a material and temperature increases, the thermal energy is converted to random kinetic energy of the molecules, and the molecules move faster. This results in an increase in the diffusion coefficient with increasing temperature. However, if we compare molecules of differing molecular weight, we find that at a given temperature a heavier molecule moves more slowly, and thus the diffusion coefficient decreases with increasing molecular weight. Example2.Molecular Diffusion Gasoline-contaminated groundwater has been transported under a house from a nearby gas station. Two meters below the dirt floor of the house's basement, the concentration of hydrocarbon vapors in the airspace within the soil is 3.10-8g/cm . Estimate the flux density of gasoline vapor transported into the basement by molecular diffusion. The diffusion coefficient for gasoline vapor in the air space within the soil is equal to cm /s. Assume that the basement is well-ventilated, so that the concentration of gasoline in the basement is very small in

42

comparison

to

the

concentration

in

the

soil.

Solution: To calculate the flux density, we use Fick's Law (equation 40), assuming that the gradient of concentration with height is linear over the 2 m depth. We will not be careful about the sign in the equation, however, since we know that the diffusive flux will be from the ground, where concentration is higher, into the basement, where the concentration is lower.

The calculated flux density of can be used to calculate the total mass flux of gasoline into the basement using equation 31. If the basement floor has an area of 100 m , then the total flux into the basement is

Turbulent Diffusion In turbulent diffusion, mass is transferred through the mixing of turbulent eddies within the fluid. This is fundamentally different from the processes which determine molecular diffusion---in turbulent diffusion; it is the random motion of the fluid that does the mixing, while in molecular diffusion it is the random motion of the pollutant molecules that is important. Earlier in this section, we used an example in which a spot of dye was dropped into the center of a river. If we follow the center of the dye spot down the river, we would see that spreading out of the spot by molecular diffusion would occur (slowly) as a result of the random motion of dye molecules across the edges of the spot. Turbulent diffusion would occur (much more quickly), as a result of eddies in the river mixing clean water from outside the spot with dye-colored water within the spot.

43

To indicate this difference in causation, the diffusion coefficient for turbulent diffusion is often referred to as the eddy diffusion coefficient. The value of the eddy diffusion coefficient depends on the properties of the fluid flow, rather than on the properties of the pollutant molecule we are interested in. Most important is the flow velocity---turbulence is only present at flow velocities above a critical level, and the degree of turbulence is correlated with velocity. (More precisely, the presence or absence of turbulence depends on the Reynolds Number, a nondimensional number which depends on velocity, width of the river or pipe, and the viscosity of the fluid.) In addition, the degree of turbulence depends on the material over which the flow is occurring, so that flow over bumpy surfaces will be more turbulent than flow over a smooth surface, and the increased turbulence will cause more rapid mixing.

Finally, the value of the eddy diffusion coefficient depends to some extent on the size scale of the problem we are considering. This is best illustrated by an example. Figure 9 shows three examples of the mixing of an isolated puff of pollutant (or spot of dye) in a turbulent flow. In Figure 9 a, the size of the pollutant puff is large compared to the size of the turbulent eddies. The result is that turbulent diffusion is slow. The opposite extreme is shown in Figure 9 b, where the size of the puff is very small compared to the turbulent eddies. In this case, the entire puff is moved along with the fluid eddies. The result is not diffusion at all---we would normally call this advection, since the puff is being ``blown'' along intact. The third example (Figure 9 c) shows the intermediate situation. Here, the
44

size of the puff is comparable to the size of the turbulent eddies, and the puff is rapidly stretched out and mixed with the surrounding fluid. In this case, the eddy diffusion coefficient would be rather large. In the real world, of course, turbulent eddies of all sizes are present simultaneously. Therefore, any given case of turbulent diffusion will be a mixture of the three situations shown in Figure 9, and only a single eddy diffusion coefficient would be used. Fick's Law applies to turbulent diffusion just as it does to molecular diffusion. Thus, flux density calculations are the same for both processes; only the magnitude of the diffusion coefficient is different. Mechanical Dispersion The final diffusion-like process that we will consider is similar to turbulence, in that it is a result of variations in the movement of the water (or air) which carries our pollutant. In mechanical dispersion, these variations are the result of (a) variations in the flow pathways taken by different fluid parcels that originate in the nearby locations near one another, or (b) variations in the speed at which fluid travels in different regions.

Dispersion in groundwater flow provides a good example of the first process. Figure 10 shows a magnified depiction of the pores within a soil sample, through which groundwater flows. (Note that, as shown in Figure 10, groundwater movement is not a result of underground rivers or creeks, but rather is caused by the flow of water through the pores of the soil, sand, or other material underground.) Because transport through the soil is limited to the pores between soil particles, each fluid particle takes a convoluted path through the soil and, as it is transported horizontally with the mean flow, it is displaced vertically a distance that depends on the exact flow path it took. The great variety of possible flow paths results in a random displacement in the
45

directions perpendicular to the mean flow path. Thus, a spot of dye introduced into the groundwater flow between points B and C in Figure 10 would be spread out, or dispersed into the region between points B and C as it flows through the soil. The second type of mechanical dispersion results from differences in flow speed, and is only important in non-steady state problems, such as the accidental, sudden release of a pollutant into a flowing stream. Anywhere that a flowing fluid contacts a stationary object, the speed at which the fluid moves will be slower near the object. For example, the speed of water flowing down a river is fastest in the center of a river, and can be very slow near the edges. Thus, if a line of dye were somehow laid across the river at one point, it would be stretched out as it flowed down the river, with the center part of the line moving faster than the edges. This type of dispersion spreads things out in the longitudinal direction---in the direction of flow---in contrast to diffusion and dispersion in groundwater, which spread things out in the direction perpendicular to the direction of mean flow. The Movement of a Particle in a Fluid---Stoke's Law In this section, we will analyze the forces which determine the movement of a particle suspended in a fluid. Suspensions of particles must be analyzed in a variety of applications, from the cleaning of particulate pollutants from coal-fired power plants and the settling out of suspended particles in wastewater treatment plants, to the dense suspension of particles in fresh cement. In each of these examples, it is useful to understand the movement of particles within the air or water fluid. The movement of a particle in a fluid can be determined by a balance of the viscous drag forces resisting the particle movement with gravitational or other forces which cause the movement. The solution to this balance of forces in the specific problem of particle settling under gravity is known as Stoke's Law. This law is derived and applied in this section. Gravitational Settling

46

Consider the settling particle shown in Figure 11. The particle is settling in a fluid, which may be air, water, or any other fluid. The particle moves in response to the gravitational force , Fg, which is equal to the mass of the particle times the gravitational constant, mpg. An upward buoyancy force is also present, and is equal to the mass of the displaced fluid times g. Let us assume that the particle is a sphere of radius Dp and density \ (g/m3), and that the fluid density is . Then the volume of the particle is , and we have

The only remaining force to determine is the drag force,FD. The drag force is the result of frictional resistance to the flow of fluid past the surface of the particle. This resistance depends on the speed at which the particle is falling through the fluid, the size of the particle, and the viscosity, or resistance to shear of the fluid. (Viscosity is essentially what one would qualitatively call the ``thickness'' of the fluid---honey has a high viscosity, while water has a relatively low viscosity and the viscosity of air is much lower yet.) Over a wide range of conditions, the friction force can be correlated with the Reynolds number. However, for most situations we are interested in, we can use Stoke's Law, which states that

where is the fluid viscosity (units of gcm-1s-1) and vr is the velocity of the particle relative to the fluid. Using this relationship, we now have an equation for all of the forces acting on the particle. The net downward force acting on the particle is equal to:

The particle will respond to this force according to Newton's second law, which states that . Thus,

47

This differential equation can be solved to determine the time-varying velocity of a particle which is initially at rest. The solution indicates that, in almost all cases of environmental interest, the period of time required before the particle reaches its final settling velocity is very short. For this reason, we will consider here only the steady-state situation in which the velocity is constant, and thus the right-hand side of equation 49 is equal to zero. In this case, Fdown=0. Setting, Fdown=0 equal to zero and noting that vr is equal to the settling velocity vs at steady state, we can rearrange equation 48 to obtain

This is the fundamental equation which is used to calculate terminal settling velocities of particles in both air and water . Because it is based on the Stoke's Law for the drag force, this equation is also often referred to as Stoke's Law, and the settling velocity is often call the Stokes velocity. A fundamental result of this equation is that the settling velocity increases as the square of the particle diameter, so that larger particles settle much faster than do smaller particles.

Example3. Settling in water. A tank called a grit chamber is to be used to remove suspended sand particles from water by allowing them to settle out. The main purpose of this chamber is to prevent the sand from wearing out pipes and pumps within the wastewater treatment plant. To design the tank in a sufficient size, we must first determine the settling velocity of the sand particles. If the sand particles are spheres with diameter Dp=100m and density p=2.65g/cm3, what would be their settling velocity? (Note that the viscosity of water is 0.01185gcm-1s-1 and the density of water is .) Solution: We will use our steady-state solution based on Stoke's Law, equation 52.

(Note that 1m=10-6m=10-4cm). The settling velocity is relatively slow. For very small particles, the drag force is significant in comparison to the gravitational force acting on their small mass.
48

Example . Settling in air. Calculate the gravitational settling velocity for two atmospheric particles having diameters of 0.1 and 100. , respectively, and with density p= 1.0g/cm3. Based on the result, determine whether removal of these particles in a settling chamber of 1 m depth would be a realistic proposition. The viscosity of air is 1.695x10-4gcm-1s-1 . The density of air is extremely small, and thus the buoyancy force may be neglected. Solution: Again, we simply need to plug the parameters of the problem into equation 52. For the 0.1 particle, we obtain

and, for the 100.

particle,

The 0.1 micron particle would take about 900 hours to settle a distance of 1 m. However, because the settling velocity is proportional to the square of the particle size, the 100 m particle settles 106 times faster, and would settle a distance of 1 m in only 3 s. It would clearly be realistic to construct a chamber with a residence time of >3 s. However, a much larger chamber would be required to reach a residence time of 9 hr, and such a chamber would likely not be economical. Note that the 100 particle in this example settles with a velocity that is much greater than that of the 100 particle settling through water in example 3.3. This is due to the much lower viscosity of air relative to water. Example . . Calculation of the minimum diameter removed by a settling chamber. A settling chamber is used to remove sand particles from the sewage flow through a wastewater treatment plant. The chamber is 2 m deep and the residence time (retention time) of water in the chamber is 4.4 hr. What is the minimum size particle which would be completely removed by settling to the bottom of the chamber during passage through the chamber? The density of sand particles is 2.65 g/cm , and the viscosity of water is 0.01185 g cm s . Assume that any particle that settles to the bottom of the chamber is removed. Solution: Since only those particles which reach the bottom of the chamber are removed, 100% removal requires that the distance settled during passage through the chamber is equal to the chamber depth. This results in a minimum settling velocity:
49

Plugging in our equation for settling velocity,

Thus, the minimum size particle removed entirely has a diameter of 13 Other applications of force balances on a particle

Analyses similar to the one we used to calculate the gravitational settling velocity can be used in a number of other applications. Essentially, in any situation where a particle moves in response to an applied force, the applied force can be balanced against the drag force to determine the particle's terminal velocity. In air pollution control, devices called electrostatic precipitators are used to remove smaller particles by applying an electric force to the particles, which causes them to move out of the airstream and onto charged collection plates. Other air pollution control devices, called cyclones, are based on the use of inertial forces to remove particles. To determine the particle removal efficiency in these devices, it is necessary to balance the drag force against a centrifugal force. Flow of water through a porous medium---Darcy's Law for groundwater flow The final problem we will consider in Part I of the course is the calculation of the rate at which water flows through a porous medium, that is, the velocity of groundwater flow. This problem is similar in some respects to the particle settling problem we just analyzed. In both situations, a fluid and a particle or particles are moving relative to one another, and the relative velocity results from a balance of the drag forces with an applied force. For a settling particle, the applied force is the downward gravitational force acting on the particle, and the drag force is due to the movement of a solid particle through a stationary fluid. In groundwater flow, the drag force is due to the movement of water (the fluid) past stationary pore surfaces. The applied force which causes groundwater movement is a pressure force. Groundwater is used extensively as a water source for agricultural and industrial use, and about half of the U.S. population uses groundwater resources for drinking water. However, groundwater flow is extremely slow in comparison to surface water flow speeds. In combination with the large volume of groundwater reservoirs, this results in slow pollutant transport but very long residence times. For this reason, contaminated
50

groundwater moves quite slowly, but once groundwater is contaminated it can be very difficult to clean up. In this section, we will present and apply the equations that govern the rate at which groundwater moves through the subsurface environment. Before we begin, however, some definitions are in order. Hydraulic Gradient The hydraulic gradient is the gradient of the height of the water table. For example, if two wells are drilled a distance of m apart and the height of the standing water within each well (called the head of water at that location) is measured, the hydraulic gradient between the two wells is given by

Just like any fluid above or below the ground, groundwater flows from regions of higher head to regions of lower head. Porosity Porosity is defined as the fraction of the total volume of soil or rock that is empty pore space capable of containing water or air. Thus, porosity, with the symbol , is defined as

Since the units cancel, porosity is a unitless value. Typical values of porosity range from 15-20% for sandstone, sand, and gravel, to 45% for clay. Darcy's Law We are now ready to consider the rate at which water flows through the subsurface environment. The force which drives this flow is proportional to the hydraulic gradient. Based on the drag force we used for particles (equation 45), we might expect that the drag force is proportional to the velocity of groundwater flow. Thus, since the driving force must be equal to the drag force, we would expect the velocity of groundwater flow to be proportional to the hydraulic gradient. Thus, we would expect

51

Where Q is the volumetric flow rate of groundwater, and A is the cross-sectional area through which the flow is occurring. This result was observed in the nineteenth century by a French civil engineer named D'Arcy. The equation he used to describe this relationship is termed Darcy's Law:

The constant K in equation 58 is called the hydraulic conductivity. The hydraulic conductivity depends on properties of the soil or rock medium and on properties of the fluid. For example, it depends on the smoothness or roughness of the pore surfaces and on how wide or narrow or how straight or tortuous the pores are. It also depends on the viscosity of the fluid (just as the drag force in Stoke's Law depends on the fluid viscosity). Representative values of hydraulic conductivity K are .2--.5 cm/s for gravel (which has big pores that are easy for water to flow through); 3x10-3,5x10-2 cm/s for sand; and 2x10-7 cm/s for clay (which is composed of very fine particles and thus has very small, tight pores that are difficult to force water through). The Darcy Velocity versus the True Velocity Darcy's Law relates the volumetric flow rate of groundwater, Q\ (units of volume water per unit time) to the cross-sectional area A\ of the soil or rock through which the flow occurs. The ratio of these two quantities gives the Darcy velocity:

However, this velocity does not reflect the true speed at which groundwater moves through the subsurface environment, because the cross-sectional area of the pores through which groundwater flows is smaller than the total soil or rock cross-sectional area. The ratio of pore cross-sectional area to total cross-sectional area is equal to the porosity, . Thus, the true velocity is given by

52

Example . . Transport time of groundwater between two wells An underground storage tank has been discovered to be leaking diesel fuel into groundwater. A drinking water well is located 200 m from the fuel spill. To ensure the safety of the drinking water supply, a monitoring well is drilled halfway between the drinking water well and the fuel spill. The difference in hydraulic head between the drinking water well and the monitoring well is 40 cm (with the head in the monitoring well higher). If the porosity is 39 percent and hydraulic conductivity is 45 m/day, how long after it reaches the monitoring well would the contaminated water reach the drinking water well? Solution: To calculate this period, we need to determine the true velocity of the groundwater between the two wells. The time for travel between the two wells will then be . The hydraulic gradient is equal to . The Darcy velocity is given by (equation 59)

The true velocity is equal to this value divided by the porosity (equation 61):

Thus, the period for flow from the monitoring well to the drinking water well is days. This result is typical of groundwater flow speeds---groundwater transport is usually very slow.

53

You might also like