Tracer Kinetic Modeling


The spatial distribution of a radiotracer in the body is time varying and depends on a number of components such as tracer delivery and extraction from the vasculature, binding to cell surface receptors, diffusion or transport into cells, metabolism, washout from the tissue, and excretion from the body. Thus the temporal component often is very important in nuclear medicine studies, and the timing of the imaging relative to the administration of the radiopharmaceutical must be carefully chosen such that the images reflect the biologic process of interest. Furthermore, the rate of change of radiotracer concentration often provides direct information on the rate of a specific biologic process. This chapter discusses how the temporal information that can be obtained from nuclear medicine studies is incorporated to provide quantitative measures of physiologic parameters, biochemical rates, or specific biologic events. Further examples are provided in reference 1.

Basic Concepts

Dynamic nuclear medicine studies enable the radiotracer concentration to be measured as a function of time, as shown in Figure 20-10 . With an understanding of the biologic fate of the radiotracer in the body, it is possible to construct mathematical models with a set of one or more parameters that can be fit to explain the observed time-activity curves. In some cases the model parameters can be related directly to physiologic or biologic quantities. Examples include tissue perfusion (measured in mL /min /g) and the rate of glucose use (measured in mol/min/g). The mathematical models that describe the time-varying distribution of radiopharmaceuticals in the body are known as tracer kinetic models.

Tracer kinetic models may be very simple. For example, one method for evaluating renal function is to measure the uptake of 99m Tc-labeled dimercaptosuccinic acid (DMSA) using a single region of interest (ROI) positioned over each kidney at one instant in time. “Function” in this case is determined in relative rather than absolute physiologic units. A more rigorous approach for evaluating kidney function is to measure glomerular filtration rates (GFRs), in mL /min, using a tracer that is filtered by the kidneys, such as 99m Tc-labeled diethylene triamine pentaacetic acid (DTPA). In this case, it is necessary to obtain serial images of the kidneys and also to collect blood samples to measure tracer concentration in the blood as a function of time. Using these data and applying an appropriate mathematical model, one can then calculate the GFR.

Each of these approaches permits an assessment of “renal function” that is based on a different model for the behavior of the kidneys. The approach of choice depends on the medical or biologic information desired, as well as on the equipment available and acceptable level of technical complexity. Developing a model requires the investigator to synthesize a large amount of biologic information into a comprehensive description of the process of interest. This chapter summarizes some of the principles and techniques in developing these models and presents some examples of tracer kinetic models currently used in nuclear medicine.

The following example illustrates the principle of tracer kinetic techniques. Figure 21-1 shows a hollow tube with a substance flowing through it. If a small amount of tracer is injected instantaneously at time t and at point A and the measured activity at point B is plotted as a function of time, the resultant time-activity curve represents a histogram of the transit times for the tracer molecules from point A to point B. If the flow rate through the tube is decreased ( dashed curve in Fig. 21-1 ), then the tracer molecules will on average take longer to get from point A to point B and the shape of the measured time-activity curve will change accordingly. This simple example illustrates conceptually how the kinetic information (i.e., the time-activity curve) varies in response to a change in a parameter in the system (flow rate). The flow rate, F, through the tube can be calculated as


F ( mL / min ) = V ( mL ) / τ ( min )

FIGURE 21-1, Illustration of use of tracer kinetics for measurement of flow. The system consists of a hollow tube characterized by flow in the direction indicated. A bolus of tracer introduced at time t and point A produces a time-activity curve at location B that depends on flow. Relatively higher flow ( solid line ) results in less dispersion and a shorter average transit time, whereas a lower flow rate ( dashed line ) produces a longer average transit time.

where V is the volume of the tube and τ is the mean transit time of the tracer molecules between points A and B. This is known as the central volume principle.

Tracers and Compartments

Most applications of tracer kinetic principles in nuclear medicine are based on compartmental models . In this section, we review the basic principles of compartmental modeling.

Definition of a Tracer

A tracer is a substance that follows (“traces”) a physiologic or biochemical process. In this chapter, tracers are assumed to be radionuclides or, more commonly, small molecules or larger biomolecules (e.g., antibodies and peptides) that are labeled with radionuclides. These labeled molecules are also known as radiotracers or radiopharmaceuticals. For simplicity, we refer to them as tracers in the remainder of this discussion. Tracers can be naturally occurring substances, analogs of natural substances (i.e., substances that mimic the natural substance), or compounds that interact with specific physiologic or biochemical processes in the body. Examples include diffusible tracers for blood flow, tracers that follow important metabolic pathways in cells, and tracers that bind to specific receptors on cell surfaces. Table 21-1 lists some examples of tracers that are used in nuclear medicine and their applications.

TABLE 21-1
Selected Examples of Tracers Used in Nuclear Medicine
Process Tracer
Blood flow/perfusion:
Diffusible (not trapped) H 2 15 O, 133 Xe, 99m Tc-teboroxime (heart)
Diffusible (trapped) 201 TlCl (heart), 99m Tc-sestamibi (heart),
13 NH 3 (heart), 82 RbCl, 99m Tc-ECD (brain),
99m Tc-tetrofosmin (heart), 62 Cu-PTSM, 99m Tc-HMPAO (brain)
Nondiffusible (trapped) 99m Tc-macroaggragated albumin (lung)
Blood volume 11 CO, 51 Cr-RBC, 99m Tc-RBC
Ventricular function 99m Tc-pertechnetate, 99m Tc-DTPA
Esophageal transit time/reflux 99m Tc-sulphur colloid
Gastric emptying 99m Tc-sulphur colloid, 111 In-DTPA
Gallbladder dynamics 99m Tc-disofenin, 99m Tc-mebrofenin
Infection 111 In-WBC, 67 Ga-citrate, 99m Tc-WBC
Lung ventilation 133 Xe, 81 Kr, 99m Tc-technegas™
Metabolism:
Oxygen 15 O 2
Oxidative 11 C-acetate
Glucose 18 F-fluorodeoxyglucose
Free fatty acids 11 C-palmitic acid, 123 I-hexadecanoic acid
Osteoblastic activity 99m Tc-MDP, 18 F
Hypoxia 18 F-fluoromisonidazole, 62 Cu-ATSM
Proliferation 18 F-fluorothymidine
Protein synthesis 11 C-leucine, 11 C-methionine
Receptor systems:
Dopaminergic 18 F-fluoro-L-dopa, 11 C-raclopride, 18 F-fluoroethylspiperone, 11 C-CFT
Benzodiazepine 18 F-flumazenil
Opiate 11 C-carfentanil
Serotonergic 11 C-altanserin
Adrenergic 123 I-mIBG
Somatostatin 111 In-octreotide
Estrogen 18 F-fluoroestradiol
ATSM, diacetyl-bis (N 4 -methylthiosemicarbazone); CFT, [N-methyl- C]-2-β-carbomethoxy-3-β-(4-fluorophenyl)-tropane; DOPA, 3,4-dihydroxyphenylalanine; DTPA, diethylenetriamine penta-acetic acid; ECD, ethyl cysteinate dimer; HMPAO, hexamethyl propylene amine oxime; MDP, methylene diphosphonate; mIBG, metaiodobenzylguanidine; PTSM, pyruvaldehyde bis(N 4 -methylthiosemithiocarbazone); RBC, red blood cell; WBC, white blood cell.

Some specific requirements for an ideal tracer include the following:

  • 1

    The behavior of the tracer should be identical or related in a known and predictable manner to that of the natural substance.

  • 2

    The mass of tracer used should not alter the underlying physiologic process being studied or should be small compared with the mass of endogenous compound being traced (a typical “rule of thumb” is that the mass of tracer should be <1% of the endogenous compound).

  • 3

    The specific activity of the tracer should be sufficiently high to permit imaging and blood or plasma activity assays without violating the first two requirements.

  • 4

    Any isotope effect (see Chapter 3 , Section B) should be negligible or at least quantitatively predictable.

If a tracer is labeled with an element not originally present in the compound (this is often the case with radionuclides such as 99m Tc, 123 I, and 18 F), it should behave similarly to the natural substance or in a way that differs in a known manner. The strictness of this requirement depends on the process under investigation. One common use of tracers in clinical nuclear medicine is to examine gross function and distribution, including blood flow, filtration, and ventilation. Although the elements represented by radionuclides such as 99m Tc, 67 Ga, 111 In, and 123 I are not normally present in biologic molecules, it is possible to incorporate these radionuclides in physiologically relevant tracers that can measure simple parameters that are related to distribution, transport, and excretion.

However, these same elements are not normally present in human biochemistry (iodine is an exception when used to study thyroid metabolism). It is therefore much more difficult to mimic a biochemical reaction sequence with these radionuclides. The biochemical systems of the body are more specific than the transport processes that move or filter fluids or gases. Biochemical systems can selectively require that compounds be of one optical polarity versus the other; that compounds fit within angstroms in the cleft of an enzyme; that chemical bond angles, lengths, and strengths are appropriate; and so forth. When a compound is labeled with a foreign species, such as 99m Tc, one cannot be sure that it will retain its natural properties and a careful examination and characterization of the compound must be undertaken. One of the advantages of radionuclides that represent elements normally involved in biochemical processes, such as 11 C, 13 N, and 15 O, is that they generally do not alter the behavior of the labeled compound.

Analog tracers are compounds that possess many of the properties of natural compounds but with differences that change the way the analog interacts with biologic systems. In many cases, analog tracers are deliberately created to simplify the analysis of a biologic system. For example, analogs that participate through only a limited number of steps in a sequence of biologic reactions have been developed in biochemistry and pharmacology. Analogs are used to decrease the number of variables that must be measured, to increase the specificity and accuracy of the measurement, or to selectively investigate a particular step in a biochemical sequence. In other cases analog tracers are used because of the need to label the tracer with an element that is not normally present in the molecule of interest. As discussed earlier, this can lead to very significant deviations in the biologic properties (particularly in small molecules) compared with the natural compound. Correction factors based on the principles of competitive substrate or enzyme kinetics are employed in studies using analog tracers to account for differences between the analog and the natural compound. A well-known and widely used example of an analog tracer in nuclear medicine is 2-deoxy-2[ 18 F]fluoro-D-glucose (FDG) to measure glucose metabolism (see Section E.5 ).

Definition of a Compartment

A compartment is a volume or space within which the tracer rapidly becomes uniformly distributed; that is, it contains no significant concentration gradients. In some cases, a compartment has an obvious physical interpretation, such as the intravascular blood pool, reactants and products in a chemical reaction, substances that are separated by membranes, and so forth. For other compartments, the physical interpretation may be less obvious, such as a tracer that may be metabolized or trapped by one of two different cell populations in an organ, thus defining the two populations of cells as separate compartments. Additionally, although the definition of a particular compartment may be appropriate for one tracer [e.g., the distribution of labeled red blood cells (RBCs) in the intravascular blood pool], it might not apply for a different tracer (e.g., the distribution of thallium or rubidium, which has both an intravascular and an extravascular distribution). Thus the number, interrelationship, organization, and definition of compartments in a compartmental model must be developed from knowledge of physiologic and biochemical principles.

Distribution Volume and Partition Coefficient

A compartment may be closed or open to a tracer. A closed compartment is one from which the tracer cannot escape, whereas an open compartment is one from which it can escape to other compartments. Whether a compartment is closed or open depends on both the compartment and the tracer. Indeed, a compartment may be open to one tracer and closed to another. If a tracer is injected into a closed compartment, such as a nondiffusible tracer in the vascular system, conservation of mass requires that after the distribution of the tracer reaches equilibrium or steady-state conditions, the amount of tracer injected, A (in becquerels or other units of activity), must equal the concentration of the tracer in the compartment, C (in Bq/mL), multiplied by the distribution volume , V d , of the compartment. Thus,


V d = A / C ( at equilibrium )

This equation is the basis for the dilution principle, which provides a convenient method for determining the distribution volume of a closed compartment, as shown by the following example.

Example 21-1

What is the distribution volume of the RBCs if 1 MBq 51 Cr-labeled RBCs is injected into the blood stream and an aliquot of blood taken after an equilibration period (10 minutes) contains 0.2 kBq/mL? Assume the hematocrit, H (fraction of the total blood volume occupied by RBCs), is 0.4.

Answer

From Equation 21-2 :


V d = ( 1000 kBq ) / ( 0.2 kBq / mL ) = 5000 mL

This result gives the total distribution volume, that is, total blood volume. The RBC volume is given by


V RBC = H × V d = 0.4 × 5000 mL = 2000 mL

More commonly, a compartment will be open; that is, the tracer will be able to escape from it. This applies, for example, to tracers that are distributed and exchanged between blood and tissue. In this case, after the tracer reaches its equilibrium distribution, * the concentration in blood will typically be different from that in the tissue ( Fig. 21-2A ). The ratio of tissue concentration C t (Bq/g) to blood concentration C b (Bq/mL) at equilibrium, is called the partition coefficient , λ, defined by

FIGURE 21-2, A, Partition coefficient, λ, for tracers that can diffuse or be transported into tissue from blood. The value of λ is given by the ratio of tissue-to-blood concentrations of the tracer when it has reached an equilibrium or steady-state condition. B, Partition coefficient also equals the ratio of the apparent distribution volume in tissue, V 1 , assuming the same tracer concentration as blood-to-tissue volume (or mass), to V t .


λ ( mL / g ) = C t ( Bq / g ) / C b ( Bq / mL )

* Note that “equilibrium” in this case means that the concentration of the tracer in the compartments has reached a constant value with time. It does not imply equilibrium in the thermodynamic sense, that is, that there is no further transport of tracer between tissue and blood. Thus tracer equilibrium is synonymous with the term steady state (see Section B.6 ).

The equilibrium blood concentration, C b , can be directly measured by taking blood samples. If one assumes that the concentration of tracer in tissue is the same as the concentration in blood ( Fig. 21-2B ), and applies Equation 21-2 , this leads to an apparent distribution volume in tissue given by V 1 = A t / C b , in which A t is the activity in the tissue. One also knows that A t = C t × V t , in which V t is the volume (or mass) of tissue; therefore combining these relationships and Equation 21-3 yields


λ = V 1 / V t

Thus another interpretation of the partition coefficient is that it is the distribution volume per unit mass of tissue for a diffusible substance or tracer. This interpretation is employed in some models for estimating blood flow and perfusion, as discussed in Section E .

Flux

Flux refers to the amount of substance that crosses a boundary or surface per unit time (e.g., mg /min or mol / min) ( Fig. 21-3 ). It also can refer to the transport of a substance between different compartments in terms of flux per unit volume or mass of tissue (e.g., mol /min /mL or mg /min /g).

FIGURE 21-3, Three-compartment system consisting of reactants in blood (R b ) and tissue (R t ) and product in tissue (P). Fluxes between the compartments, indicated by arrows, are products of the first-order rate constants and the respective compartmental concentrations.

Flux is a general term that can refer to a variety of processes. For example, the total mass of RBCs moving through a blood vessel per unit time is a flux. The “boundary” or “surface” in this case could be any transverse plane through the vessel. The amount of glucose moving across a cell membrane per unit time also is a flux. Fluxes therefore may either be closely related or unrelated to blood flow.

Rate Constants

Rate constants describe the relationships between the concentrations and fluxes of a substance between two compartments. For simple first-order processes, the rate constant, k, multiplied by the amount (or concentration) of a substance in a compartment determines the flux:


flux = k × amount of substance in compartment

For first-order processes, the units of k are (time) –1 . If “amount” refers to the mass of tracer in the compartment, the units of flux are mass/time (e.g., mg/min). If “amount” refers to concentration of tracer in the compartment, the units of flux are mass/time per unit of compartment volume (e.g., mg/min/mL), or mass/time per unit of compartment mass (e.g., mg/min/g). Note that, as illustrated by Figure 21-3 , different directions of transport between two compartments can be characterized by different rate constants.

A first-order rate constant also may represent the fractional rate of transport of a substance from a compartment per unit time. For example, a rate constant of 0.1 min –1 corresponds to a transport of 10% of the substance from the compartment per minute. The inverse of the rate constant, 1/ k , is sometimes referred to as the turnover time, or mean transit time , τ, of the tracer in the compartment (in this example, 10 minutes). Similarly, the half-time of turnover, t 1/2 , that is, the time required for the original amount of tracer in the compartment to decrease by 50% (assuming no back transfer into the compartment), is given by


t 1 / 2 = 0.693 / k

Thus the fractional rate constant k is analogous to the decay constant λ for radioactive decay, whereas the mean transit time is analogous to the average lifetime of a radionuclide (see Chapter 4 , Section B.3). In first-order models, transport out of a compartment through a single pathway (without back-transport) is described by a single exponential function, e kt , analogous to the radioactive decay factor e –λ t .

If there is more than one potential pathway for a tracer to leave a compartment, each characterized by a separate rate constant, k i , then the turnover time of the tracer in the compartment is the inverse of the sum of all these rate constants and the half-time of turnover is


t 1 / 2 = 0.693 / ( k 1 + k 2 + + k m )

where m is the number of pathways by which the tracer can leave the compartment.

Most compartment models used in nuclear medicine are based on the assumption that first-order kinetics describe the dynamics of the system of interest. The tracer kinetics of such systems are linear. That is, doubling the input (amount or concentration) doubles the output (flux) of the system. As shown in Section E , linear first-order tracer kinetic models adequately describe many systems even when the dynamics of the natural substances are nonlinear.

A more general expression for the relationship among rate constants, fluxes, and concentrations (or masses) is


flux = k × ( mass or concentration of substance ) n

where n refers to the order of the reaction. The units of rate constants for n th order reactions (in terms of concentration) are [concentrations (1— n ) • time –1 ]. Thus only first-order rate constants represent a constant fractional turnover and Equations 21-6 and 21-7 apply only to first-order processes.

Figure 21-3 illustrates a three-compartment system consisting of a blood compartment separated by a membrane barrier (e.g., capillary wall) from two sequential tissue compartments. R and P refer to chemical reactant and product, whereas the subscripts b and t refer to reactant in blood and tissue compartments, respectively. [R b ], * [R t ], and [P] are the blood and tissue concentrations of reactant and product, whereas the fluxes between the compartments are the first-order rate constants, k 1 , k 2 , k 3 , and k 4 , multiplied by corresponding concentrations. The thicknesses of the arrows in Figure 21-3 are proportional to the magnitude of the corresponding rate constant. In this example, the rate constants into and out of tissue are larger than the corresponding rate constants between the reactant and product compartments in tissue. Thus the majority of the reactant initially transported into the tissue space is transported back into blood without undergoing any biochemical reactions. This is a common occurrence in actual biochemical systems and introduces a reserve capacity into the system that can accommodate changes in metabolic supply and demand (e.g., by changing k 3 ).

* The notation [X] is used to denote the concentration (usually in units of g/mL or mol/mL) of substance X.

Figure 21-4 illustrates the relationship between first-order rate constants and the relative concentrations of the substrates in a biochemical sequence. If a substrate (S) and enzyme (E) combine to form a substrate-enzyme complex (SE), which then dissociates into a product (P) with release of the enzyme, the fluxes of the first-order reaction steps are concentrations multiplied by the corresponding rate constants. If a small amount of labeled substrate is introduced into the system at time zero, the tracer will go through the reaction steps, producing concentrations of labeled S, SE, and P as shown in the graphs in Figure 21-4 . If k 3 (the forward rate constant for the reaction converting SE to E and P) is reduced by 50% with all the other rate constants remaining unchanged, the concentrations of labeled S, SE, and P are then represented by the dotted orange lines in Figure 21-4 . Decreasing k 3 causes a slower production of P and causes a compensatory increase in labeled S and SE.

FIGURE 21-4, Illustration of tracer kinetics of a chemical reaction sequence. First-order rate constants ( k 0 , k 1 , k 2 , k 3 , k 4 , k 5 ) characterize the various reaction steps, whereas S refers to substrate, E refers to enzyme, P is product, and SE is the substrate-enzyme complex. The time-activity relationships for concentrations of labeled S, SE, and P are shown for a particular value of k 3 ( solid purple line ) and with k 3 reduced by 50% ( dotted orange line ).

You're Reading a Preview

Become a Clinical Tree membership for Full access and enjoy Unlimited articles

Become membership

If you are a member. Log in here