Pharmacologic Principles


CHAPTER 4 Pharmacologic Principles







DEFINITIONS IN PHARMACOKINETICS


Pharmacokinetic information is used to determine drug dosage regimens in clinical patients. An understanding of the way in which drug dosage regimens are derived and how they can be adjusted for different disease states requires knowledge of some basic pharmacokinetic terms. Mathematical models provide equations to describe drug concentration as a function of time. With an open model the drug is eliminated from the body. An open model describes the fate of most drugs. With a closed model the drug is recirculated within the body (e.g., a drug that undergoes enterohepatic recirculation). In pharmacokinetic models the body is represented by a series of compartments that communicate reversibly with one another. A compartment is a tissue or group of tissues with similar blood flow and drug affinity. A drug is assumed to be uniformly distributed within a compartment and can move dynamically in and out of compartments. Rate constants represent the entry and exit of drugs from each compartment. The central compartment is made up of the highly perfused tissues that equilibrate rapidly with the drug. Overall drug elimination occurs mainly from the central compartment, because the kidneys and liver are well-perfused tissues. The peripheral compartment is made up of less-perfused tissues such as muscle and connective tissues. The deep compartment consists of slowly perfused tissues or depot tissues such as fat and bone. The presence of a deep compartment for drug distribution is important for toxins and drug residues. Most drugs in clinical use are described by one or two compartment models. Models with more than three compartments are usually not physiologically relevant. Describing drug disposition with compartment models creates differential equations that describe drug concentration changes in each compartment and provides a visual representation of the rate processes among compartments.



RATES AND ORDERS OF REACTIONS


The drug absorption or elimination rate is the speed with which it occurs. If the amount of drug in the body is decreasing over time, then the elimination rate is expressed as follows:



image



The absorption and elimination rate of a drug is determined experimentally by measuring the plasma drug concentration at given time intervals. Rate constants relate the observed rate of a kinetic process to the drug concentration that controls the process. The elimination rate constant (K) is equal to the rate of drug elimination divided by the amount of drug in the body. The absorption rate constant (Ka) describes the rate of drug absorption into the central compartment. Reaction order refers to the way that drug concentration influences reaction rate.


With a zero order reaction the amount of drug changes at a constant time interval, regardless of the drug concentration. The rate of drug elimination is as follows:



image



where K0 is the zero order rate constant in mg/ml min. A graph of drug concentration versus time on regular graph paper for a zero order reaction produces a straight line (Figure 4-1), described by the equation:




image



where C is the drug concentration at any time t, and C0 is the drug concentration at time zero. For most drugs zero order elimination occurs only when elimination mechanisms become saturated. Renal tubular secretion and bile secretion of drugs are examples of potentially saturable processes. The most well-known zero order reaction is the oxidation of ethanol in humans. The alcohol dehydrogenase system becomes saturated with very small amounts of ethanol. To achieve mild intoxication (1 mg/ml) throughout a 75-kg person, an intake of about 56 ml of absolute alcohol (or 4 oz of whiskey, vodka, or gin) is required. The maximum amount of alcohol that can be eliminated is 10 ml per hour; therefore it takes 5 hours to totally eliminate the original 56 ml. Therefore to maintain a constant level of mild intoxication requires only 10 ml of ethanol or 25 ml of liquor per hour. In veterinary medicine, drugs with well-known zero order elimination include phenylbutazone in horses and deracoxibin dogs. Once elimination processes are saturated, increased dosages of such drugs result in wildly unpredictable plasma concentrations and easily result in toxicity.


With a first order reaction, the amount of drug changes at a rate proportional to the amount of drug remaining. The first order elimination rate is expressed as follows:



image



where K is the first order rate constant, is expressed in units of time-1 (min-1 or hr-1), and defines the fraction of drug eliminated from the body per unit time; C is the plasma drug concentration at any time t. Although K remains constant, the rate (>C/>t) is always changing because C is always decreasing. A graph of drug concentration versus time for a first order reaction produces an exponential curve on regular graph paper but produces a straight line on semilogrithmic graph paper (Figure 4-2) and is described by the following equation:




image



where C is drug concentration at any time t, K is the first order rate constant in minutes or hours, and C0 is the drug concentration at time zero (the moment of injection). Most drugs are absorbed and eliminated by first order processes. Glomerular filtration by the kidney is a first order process.



CLINICAL APPLICATION OF COMPARTMENTAL MODELING, RATES, AND ORDERS OF REACTIONS


The aforementioned concepts can be combined to mathematically describe the changes in the drug concentration in the body over time. Drug disposition described by a one-compartment open model with intravenous (IV) injection and first order elimination (Figure 4-3) means that the body acts as one homogeneous compartment. A drug’s concentration in one part of the body is assumed to be proportional to its concentration in any other part. Many drugs administered by routes other than IV, such as oral, subcutaneous, intramuscular, or intradermal, are described by a one-compartment open model with first order absorption (Ka) and elimination (K) (Figure 4-4). With a two-compartment open model with IV injection and first order elimination, the model assumes the body acts as two compartments: the central compartment (blood and highly vascularized tissues) and a peripheral compartment (less vascularized tissues). Most drugs administered in veterinary medicine are described by this model (Figure 4-5). Elimination is considered to occur only from the central compartment, because the liver and kidneys are highly vascularized tissues. The plasma concentration versus time graph does not produce a straight line on semilogrithmic paper but can be broken into two sections and described by the following biexponential equation:






image



Where C is the concentration at any time t, A is the y-intercept of the first portion of the curve extrapolated to zero, and α is the slope of the line; B is the y-intercept of the latter portion of the curve extrapolated to zero, and β is its slope. The movement of drug between the central and peripheral compartments is described by the rate constants K12 and K21.


For some concentration versus time data, the line can be broken into three or more straight lines, and described mathematically with three or more exponential terms. Theoretically, drug distribution in the body can be described by as many compartments as there are different tissues, but for practical purposes more than three compartments models are not necessary. Drugs that are described by three compartment models usually have some tissue site where the drug is sequestered and slowly eliminated from the body, such as the aminoglycosides, which sequester in the renal tubular epithelial cells, and oxytetracycline, which sequesters in teeth and bone.



DISTRIBUTION OF DRUGS IN THE BODY


The volume of distribution (Vd) of a drug is the mathematical term used to describe the apparent volume of the body in which a drug is dissolved.1 The Vd is the parameter used to assess the amount of drug in the body from the measurement of a “snapshot” plasma concentration. The numerical value of Vd can give some indication of the distribution of the drug in the body. A drug’s distribution is determined by its ability to cross biologic membranes and reach tissues outside the vascular system. The physical characteristics of the drug molecule, such as ionization, lipid solubility, molecular size, and degree of protein binding, determine its ability to cross biologic membranes.


Three volumes of distribution (Vd) are reported in the veterinary literature: the volume of the central compartment, the steady-state volume of distribution, and the volume of distribution calculated by the area method. Conceptually, the easiest demonstration of the volume of distribution is with the volume of the central compartment (Vdc). Just after an IV dose, plasma drug concentration is maximal (Figure 4-6). Assuming that the instant drug concentration (C0) results from the drug mixing in the blood, the Vdc is the apparent volume from which drug elimination occurs, because the kidneys and liver belong to the central compartment, and is calculated from the following equation:




image



where C0 is the concentration at time zero, extrapolated from the plasma concentration versus time graph. To understand what the Vdc for a drug represents, consider the body as a beaker filled with fluid (Figure 4-7). The fluid represents the plasma and other components of extracellular water. If a drug is administered intravenously, it rapidly distributes in the extracellular fluid. If the drug does not readily cross lipid membranes, it will be confined mainly to the extracellular fluid and a plasma sample therefore will have a high drug concentration. The higher the measured concentration in relation to the original dose, the lower the numerical value for Vdc. Drugs such as the β-lactam and aminoglycoside antibiotics are poorly lipid soluble and therefore remain predominantly in the extracellular fluid and have low values for Vd. In contrast, some drugs readily cross lipid membranes and distribute into tissues. This is represented by the beaker on the right, where the stars at the bottom of the beaker represent drug molecules that have been taken up by tissues. A plasma sample will have a low drug concentration in proportion to the original dose and therefore will have a high numerical value for Vdc. Given the limitations on measuring drug concentrations at “time zero” and using the aforementioned formula, the measured concentration of highly lipid-soluble drugs can be low enough to result in a value of Vdc that is greater than 1 L/kg, so it is often referred to as an apparent volume of distribution. In the example on the right, the “apparent” Vd is 2 L/kg even though the beaker contains only 1 L of fluid.



For most drugs after a single IV dose, the drug is distributed and begins to be eliminated simultaneously. When concentrations are measured and the data graphed, there is a distribution phase, wherein the plasma drug concentration that is due to elimination and not distribution increases until it reaches an asymptotic value at which pseudoequilibrium is achieved (see Figure 4-6). When pseudoequilibrium is reached, the movement of drug between the peripheral and central compartments reaches equilibrium, and decreasing plasma concentrations are now due only to irreversible elimination (described by the elimination rate constant, β). The applicable Vd value in this situation is the volume of distribution by area (Vdarea),



image



where AUC0-∞ is the area under the plasma concentration time curve from zero to infinity. To be calculated accurately, the amount of drug that enters the systemic circulation must be accurately known and the terminal phase must be a pure elimination phase. An inaccurate Vdarea is frequently published for “long-acting” intramuscular or subcutaneous administered drugs, where prolonged elimination is due to delayed absorption.


With an IV infusion or with a multiple dose regimen, the rate of drug entry into the body is equal to its elimination rate, and the body becomes a closed system with no clearance. In this situation the correct Vd to describe distribution is the Vd at steady-state (Vdss; Figure 4-8):




image




Clinical Use of the Different Volume of Distribution Values


The Vdc is used to predict the initial plasma drug concentration after an IV bolus of a drug when a loading dose is needed to rapidly achieve a therapeutic drug concentration. The Vdss is used to calculate a loading dose when it is clinically necessary to rapidly reach steady-state concentrations. The Vdarea is used to predict the amount of drug remaining in the body. For all drugs the value of Vdarea is greater than Vdss, but generally the difference is small and the values are used interchangeably. However, with the IV administration of drugs that are rapidly eliminated into urine (e.g, aminoglycosides), Vdarea can be much larger than Vdss because a large fraction of the drug is eliminated before pseudoequilibrium is reached.


It is useful to compare a drug’s Vd to the distribution of water in the body in order to get an idea of its distribution. Drugs with a Vd value of less than 0.3 L/kg are predominantly confined to the ECF, whereas drugs with a Vd value of greater than 1 L/kg are highly lipid soluble and tend to distribute out of the extracellular fluid and into tissue compartments (Box 4-1). Although the value of Vd does not confirm penetration of a drug into specific tissues, in general the higher the value of the Vd, the more likely it is that the drug will reach sequestered sites such as the brain and cerebrospinal fluid, the prostate and other sex organs, the eye, and the mammary gland. Studies must be performed to confirm that therapeutic concentrations are achieved in such sites.





Bioavailability


Bioavailability (F) is a measure of the systemic availability of a drug administered by a route other than IV.3 Bioavailability is determined by comparing the area under the plasma drug concentration curve versus time (AUC) for the extravascular formulation to the AUC for the IV formulation. The AUC is calculated by computer or by the trapezoidal method, wherein the entire curve is divided into trapezoids, then the area of each trapezoid is calculated and summed to give the AUC. For an orally administered drug, use the following equation:



image



If F is significantly less than 100%, the drug dose must be increased to achieve systemic drug concentrations similar to the IV formulation:



image



If the oral formulation of a drug has a mean bioavailability of 50%, the drug dose must be doubled to achieve the same concentrations in plasma as achieved using the IV formulation. However, the variability of the bioavailability in the population is more clinically significant than the mean. To make sure that the horse with the poorest absorption is dosed appropriately, the dose must be increased according to the lowest bioavailability, not the mean. For example, if a drug has a mean F of 50% with a range of 20% to 70%, then to achieve an exposure of 100% for all the treated horses, the dose must be multiplied by 5, not just 2. However, if this is done, the horses with an F of 70% will be overdosed by a factor of 3.5. For a drug with a narrow therapeutic window and a poor bioavailability, there may be no dose that is ideal for all horses in the population. Low bioavailability of antimicrobials and anthelmintics is a major cause of subtherapeutic dosages that promote drug resistance. Poor oral bioavailability is a major limitation of many drugs administered to horses.4



Lipid Solubility and Drug Ionization (the pH-Partition Hypothesis)


The degree of lipid solubility determines how readily a drug will cross biologic membranes. Drugs are classified as lipid soluble (or nonpolar) versus water soluble (or polar). Highly lipophilic drugs diffuse easily across almost all tissue membranes. Most of the drugs used in equine practice exist as weak acids or weak bases. Their lipid solubility depends a great deal on their degree of ionization (charged state). An ionized drug is hydrophillic and poorly lipid soluble. A nonionized drug is lipophilic and can cross biologic membranes. The degree of ionization for a weak acid or weak base depends on the pKa of the drug and the pH of the surrounding fluid. At a given pH, there is an equilibrium between the ionized and nonionized proportions of drug. When the pH is equal to the pKa of the drug, then the drug will be 50% ionized and 50% nonionized (log 1 = 0). As the pH changes, the proportion of ionized to nonionized drug will change according to the Henderson-Hasselbach equations:


For a weak acid:



image



For a weak base:



image



Whereas the precise ratios of ionized versus nonionized drug can be calculated from the Henderson-Hasselbach equations, the relevance of the equations can be understood by simply remembering the sentence “like is nonionized in like.” For example, a weak acid will be most nonionized in an acidic environment, so aspirin is most nonionized in the stomach and is readily absorbed. The fluid of most sequestered sites in the body (cerebrospinal fluid, accessory sex gland fluid, milk, abscesses) has a pH more acidic than plasma. In cattle with mastitis weak acid antibiotics are typically administered by intramammary infusion, whereas weak bases are administered parenterally. This makes sense according to the pH-partition concept. Weak bases in the plasma are highly nonionized and readily cross into the mammary gland. Then, as the equilibrium shifts, they become “ion-trapped” in the more acidic milk, but the fraction of nonionized drug in the mammary gland is available to cross the bacterial cell membrane for antimicrobial action. Weak acids such as penicillins and cephalosporins are highly ionized in plasma and therefore do not penetrate into the mammary gland very well, so these are most effective when administered by local infusion into the udder, where the extremely high local concentrations negate local pH effects.


Typically, drugs that are weak acids will have low Vd values and weak bases will have high values for Vd (Box 4-2). Amphoteric drugs such as the fluoroquinolones and tetracyclines have acidic and basic groups on their chemical structures. These drugs have a pH range where they are maximally nonionized. For example, enrofloxacin is most lipid soluble (nonionized) in the pH range of 6 to 8, so it is lipid soluble at most physiologic pHs. In acidic urine significant ionization occurs, which reduces enrofloxacin’s antibacterial activity. But this reduction in activity is offset by the extremely high concentrations of enrofloxacin achieved in urine, so it is of no clinical importance. Despite being weak bases, the aminoglycosides are very large, hydrophilic molecules and have high pKa values, so they are highly ionized at physiologic pHs. Therefore parenterally administered aminoglycosides do not cross lipid membranes well and do not achieve therapeutic concentrations in milk, accessory sex gland fluids, abscesses, or cerebrospinal fluid.




Drug Protein Binding


Protein binding can involve plasma proteins, extracellular tissue proteins, or intracellular tissue proteins. Many drugs in circulation are bound to plasma proteins, and because bound drug is too large to pass through biologic membranes, only free drug is available for delivery to the tissues and to produce the desired pharmacologic action. Therefore the degree of protein binding can greatly affect the pharmacokinetics of drugs. Acidic drugs such as nonsteroidal anti-inflammatory drugs tend to bind predominantly to albumin.5 Albumin is the most abundant plasma protein, and it is critical to maintaining the colloidal oncotic pressure in the vascular system. As a negative acute phase protein, albumin concentration decreases during inflammation. Hypoalbuminemia results from decreased production, seen with severe hepatic insufficiency, or by loss through increased rates of urinary excretion, such as in nephrotic syndrome or with mucosal damage, as with protein losing enteropathies. Basic drugs typically bind to α-1 acid glycoprotein, which is an acute phase protein, whose hepatic production increases significantly with inflammatory conditions.6 Other proteins, including corticosteroid binding globulin, are important for binding of some specific drugs but are less important in overall drug-protein binding.7 There is equilibrium between free and bound drug, however, just like the relationship of ionized and nonionized drug molecules. Protein binding is most clinically significant for antimicrobial therapy, where a high degree of protein binding serves as a drug “depot,” allowing for increased duration of the time the drug concentration remains above the bacterial minimum inhibitory concentration, adding to antimicrobial efficacy.8 For other drugs changes in plasma protein binding can influence individual pharmacokinetic parameters, but changes in plasma protein binding usually do not influence the clinical exposure of the patient to a drug. Changes in protein binding caused by drug interactions are assumed to instantaneously change free drug concentrations and have been frequently cited as a cause of adverse drug reactions. But the increase in free drug concentration is only transient, because drug distribution and drug elimination change to compensate. The often-cited example of the concurrent administration of phenylbutazone and warfarin leading to bleeding caused by increased free concentrations of warfarin is erroneous. The true interaction is from phenylbutazone-induced inhibition of the hepatic metabolism of warfarin, which results in increased plasma concentrations and increased anticoagulant effect.7 Therefore adjustments in dosing regimens because of hypoproteinemia or concurrent administration of highly bound drugs are not necessary except in the rare case of a drug with a high hepatic extraction ratio and narrow therapeutic index that is given parenterally (e.g., IV dosing of lidocaine).9




ELIMINATION RATE CONSTANT AND ELIMINATION HALF-LIFE


The rate of elimination for most drugs is a first order process. The elimination rate constant (K) represents the sum of drug elimination by excretion and metabolism. Drug elimination is considered always to occur from the central compartment, because the liver and kidney are well-perfused tissues. The elimination rate constant is used to calculate the drug’s half-life (T½), or the time required for drug concentration to decrease by one half (Figure 4-9). For first order reactions, T½ is constant across the plasma concentration versus time curve and is calculated from:




image



where 0.693 = ln2 (the natural logarithm of 2). Mean residence time (MRT) is roughly the equivalent of T½ when pharmacokinetics are calculated using statistical moment theory. The MRT is the time it takes for drug concentration to decrease by 63.2%, so the MRT value is typically slightly greater than T½. The T½ determines the drug dosage interval, how long a toxic or pharmacologic effect will persist, and drug withdrawal times for food animals or performance horses. Notice that it takes 10 T½s to decrease the plasma concentration by 99.9% (Table 4-1). Knowing a drug’s plasma T½ can give the clinician some idea of the drug’s withdrawal time for food or performance animals. However, for drugs that undergo hepatic metabolism (e.g., phenylbutazine) or drugs that sequester in specific tissues (e.g., aminoglycosides, isoxuprine), simply multiplying the T½ by a factor of 10 for a withdrawal time may not be sufficient to prevent violative residues. Also note that doubling a drug dose does not double the withdrawal time; it merely adds one half-life to the time it takes to reach the acceptable threshold concentration (Figure 4-10).


TABLE 4-1 Half-Life of Elimination of a Drug







































Number of Half-Lives Fraction of Drug Remaining (%)
0 100
1 50
2 25
3 12.5
4 6.25
5 3.125
6 1.56
7 0.78
8 0.39
9 0.195
10 0.0975




Clearance


Clearance is a measure of drug elimination from the body without reference to the mechanism of elimination. It is always reported in pharmacokinetic papers, but its significance is rarely explained in pharmacokinetic studies in horses. Clearance is the most important pharmacokinetic parameter because it is the only parameter that controls overall drug exposure and it is used to calculate the dosage required to maintain a specific average steady-state concentration.10


Clearance (Cl) is the total drug clearance and is the sum of renal clearance (ClR), hepatic clearance (ClH), and all other elimination mechanisms. By definition, Cl is the volume of fluid containing drug that is cleared of drug per unit of time (ml/kg/min). The most frequent technique for determining plasma Cl is to administer a single IV dose of a drug and then measure plasma concentrations over time. Then,



image



where AUC is the area under the plasma concentration time curve.


If the body is considered as the whole system clearing the drug, Cl can also be determined by the animal’s cardiac output and the extraction ratio (E), where E is a numerical value between 0 and 1 that is the percentage of the drug that is cleared by a single pass through the clearing organ:



image



For a drug with an extraction ratio of 1 (100% removal by the liver and kidney on the first pass), then the expected value of Cl is about 50% of the cardiac output, because blood flow to the liver and the kidneys represents about half of the cardiac output.


In contrast to Vd values, Cl values have to be interpreted according to the value of cardiac output for the species involved. Given that most drugs are extracted primarily by renal and hepatic mechanisms, the extraction ratio is considered high if E is greater than 0.7, medium if E equals 0.3, and low if E is less than 0.1. Because the liver and kidneys receive about 50% of cardiac output, then the overall E is high if it is greater than 0.35, medium if it is about 0.15, and low if it is less than 0.05. From this and the cardiac output of the species, breakpoint values can be determined to classify drugs as having high, medium, and low clearance. For the horse with a cardiac output of 55 ml/kg/min, a high Cl value is 19 ml/min/kg, medium Cl is 8.25 ml/min/kg, and low Cl is 3.6 ml/min/kg.


It is sometimes difficult to understand the difference between the elimination half-life and clearance. The relationship is as follows:



image



Consider the values for clearance and T½ values for four antimicrobial drugs (Table 4-2). Note that the plasma clearance values are similar, but the elimination half-lives are very different. Because the T½ is influenced by the extent of drug distribution, the drugs have similar clearance, but oxytetracycline has the largest Vd and the longest T½. Because T½ is derived from rate constants and does not have a physiologic basis, it is influenced by the sensitivity of the analytical method and by many pharmacokinetic parameters, and it is a poor parameter alone to evaluate physiologic (e.g., age, sex) or pathologic (e.g., renal failure) changes that effect drug disposition.




Renal Clearance of Drugs


Renal excretion is the major route of elimination from the body for most drugs. Drug disposition by the kidneys includes glomerular filtration, active tubular secretion, and tubular reabsorption (Figure 4-12), such that renal drug clearance is defined by the following equation:




image






Glomerular filtration (ClF) occurs with small molecules (<300 molecular weight) of free drug (not bound to plasma proteins). Large molecules or protein-bound drugs do not get filtered at the glomerulus because of size and electrical hindrance. The kidneys receive approximately 25% of cardiac output, so the major driving force for glomerular filtration is the hydrostatic pressure within the glomerular capillaries. Glomerular filtration rate (GFR) is estimated by measuring a substance or drug that is eliminated only by glomerular filtration, such as creatinine or inulin.


If ClR is greater than ClF, then some tubular secretion is occurring. Active tubular secretion is a carrier-mediated transport system, located in the proximal renal tubule. It requires energy input because the drug is moved against a concentration gradient. Two active tubular secretion systems have been identified: anion secretion for acids and cation secretion for bases. Drugs with similar structures may compete with each other for the same transport system. For example, probenecid competes with penicillin or the fluoroquinolones for the same transport system, effectively decreasing Cl of these antimicrobials. In patients with reduced functional renal tissue, remaining transport systems become easily saturated and drug accumulation occurs.


If ClR is less than GFR, tubular reabsorption of drug is occurring. Tubular reabsorption is an active process for endogenous compounds (e.g., vitamins, electrolytes, glucose). It is a passive process for the majority of drugs. It occurs along the entire nephron but primarily in the distal renal tubule. Factors that affect reabsorption include the pKa of the drug, urine pH, lipid solubility, drug size, and urine flow. Drug reabsorption is highly dependent on ionization, which is determined by the drug’s pKa and the pH of the urine. According to the Henderson-Hasselbach equation, a drug that is a weak base will be mainly nonionized in alkaline urine and a weak acid will be mainly ionized in alkaline urine. The nonionized form of the drug is more lipid soluble and has greater reabsorption (Table 4-3). The pKa of a drug is constant, but urinary pH is highly variable in animals and varies with the diet, drug intake, time of day, and systemic acidosis/alkalosis. Species differences have a major influence on the renal excretion of ionized drugs. Carnivores, with a urine pH of 5.5 to 7.0, will have a greater renal excretion of basic drugs than herbivores, with a urine pH of 7.0 to 8.0.





Biotransformation (Hepatic Metabolism) of Drugs


Metabolism is necessary for removal of lipophilic drugs from the body (Figure 4-13). Biotransformation depends on the chemical composition of the liver, activity of major drug metabolism enzymes, hepatic volume (perfusion rate), drug accessibility to and extraction by hepatic metabolic sites, and physicochemical properties of the drug. Biotransformation of a parent drug results in metabolites that may be active or inactive themselves. A prodrug is a drug administered in an inactive form that must be biotransformed to its active form, such as prednisone to prednisolone. Drug metabolic pathways are divided into phase I and phase II reactions. Phase I reactions (oxidation, reduction, hydrolysis, hydration, dethioacetylation, isomerization) typically add functional groups to the drug molecule necessary for phase II reactions. Phase II reactions (glucuronidation, glucosidation, sulfation, methylation, acetylation, amino acid conjugation, glutathione conjugation, and fatty acid conjugation) typically include conjugation reactions that increase the water solubility of the drug, facilitating excretion from the body.



Among the reactions catalyzed by drug metabolism enzymes, the cytochrome P450 mixed function oxidase system is the most intensively studied. This reaction catalyzes the hydroxylation of hundreds of structurally diverse drugs, whose only common characteristic is high lipid solubility. Species differences in drug metabolic rate are the primary source of variation in drug activity and toxicity. Cats have a poor ability to glucuronidate drugs, pigs are deficient in sulfate conjugation, and dogs are relatively poor acetylators.



Induction and Inhibition of Metabolism


Metabolism of drugs can be substantially affected by enzyme induction or inhibition by other drugs or chemicals (Box 4-3). In some cases the drug itself may alter its own metabolic fate by induction or inhibition. Many drugs are capable of inducing enzyme activity, thereby increasing the rate of metabolism and hepatic clearance of concurrently administered drugs, which typically results in a decreased pharmacologic effect. Enzyme induction typically occurs slowly, requiring several weeks to reach maximum effect. Induction is accompanied by increased hepatic ribonucleic acid (RNA) and protein synthesis and increased hepatic weight. Enzyme induction is important in the pathogenesis of hepatotoxicity and therapeutic failure of many drugs. Phenobarbital is a potent enzyme inducer known for hepatotoxicity and for inducing its own metabolism. Rifampin induces the metabolism of azole antifungals; concurrent administration with itraconazole results in subtherapeutic itraconazole concentrations.



Drug-induced enzyme inhibition also occurs and typically results in prolonged clearance of a concurrently administered drug. The potential for toxicity or an exaggerated pharmacologic response is increased. In contrast to induction, inhibition occurs rapidly. Erythromycin and enrofloxacin are known inhibitors of the metabolism of theophylline; concurrent administration can cause central nervous system toxicity and seizures.11,12




DRUG ACCUMULATION


Drugs are often given in multiple-dosage regimens. To predict plasma drug concentrations, it is necessary to decide whether successive doses of a drug have any effect on the previous dose. The principle of superposition assumes that early doses of drug do not affect the pharmacokinetics of subsequent doses. For most drugs, as equal doses are given at a constant dosage interval, the plasma concentration-time curve plateaus and a steady-state is reached. At steady-state the plasma drug concentration fluctuates between a maximum concentration (Cmax, or peak) and minimum concentration (Cmin, or trough). Once steady-state is reached, Cmax and Cmin are constant and remain unchanged from dose to dose (Figure 4-14). The time to steady-state depends solely on the elimination half-life. It takes approximately six T½s to reach 99% steady-state levels. The drug dose and dosage frequency influence the values of Cmax and Cmin at steady-state, while the dosage frequency and T½ influence the fluctuation between Cmax and Cmin.




Clinical Consequences of Dosage Intervals Less than the Half-Life


Drugs such as phenobarbital, potassium bromide, phenylbutazone, and digoxin are commonly given at dosage intervals much shorter than their T½=s (Figure 4-15). This results in a Cmax at steady-state that is greater than the peak concentration after a single dose. There is minimal fluctuation between Cmax and Cmin, and missing a single dose will not affect plasma concentrations greatly. There is a lag time to reach the desired plasma concentrations at steady-state, and there will be a lag time for plasma concentrations to change in response to a dose change.




Clinical Consequences of Dosage Intervals Greater than the Half-Life


Drugs such as IV formulations of penicillin and cephalosporins are administered at dosage intervals that are greater than the T½ (Figure 4-16). As the dosage interval increases, Cmax at steady-state is closer in value to the peak concentration after a single dose. If the dosage interval is greater than 10 T½ s (the time required to eliminate 99.9% of the previous dose), drug accumulation essentially does not occur. There is marked fluctuation between Cmax and Cmin (peak and trough), and missing a dose will greatly affect plasma concentrations. However, there is minimal lag time to achieve the desired plasma concentration.




DESIGNING DRUG DOSAGE REGIMENS


The success of drug therapy is highly dependent on the dosage regimen design. Not all drugs require rigid individualization of the dosage regimen. In the case of antimicrobials with a broad safety range, such as the penicillins and cephalosporins, the dosage is not titrated precisely but rather determined on the basis of clinical judgment to maintain an effective plasma concentration above the minimum inhibitory concentration of the bacterial pathogen. For drugs with a narrow therapeutic margin, such as digoxin, the aminoglycosides, and theophylline, the individualization of the dosage regimen is very important. The objective of the dosage regimen for these drugs is to produce a safe plasma drug concentration that does not exceed the minimum toxic concentration or fall below a critical minimum concentration below which the drug is not effective. Factors that influence the concentration of drug attained at the drug’s site of action include the dose administered, the route of administration, release and absorption of drug from dosage form, the extent of drug distribution, and the rate of drug elimination.


The horse’s age may have a profound effect on drug disposition (Table 4-4).13 The definition of geriatric varies between species, and in small animals it varies between breeds. Body composition and regional blood flow change in geriatric horses. Cardiac output decreases, so regional and organ blood flow also decrease. These changes have an impact on drug absorption, distribution, and elimination. Blood flow is preferably redistributed to the brain and heart, so there is an increase in risk of drug toxicity in these organs. Gastrointestinal motility and absorptive capacity are reduced. Hepatocyte number and function decrease along with hepatic and splanchnic blood flow. As renal blood flow decreases, GFR and active secretory capacity of the nephron decreases, resulting in decreased renal clearance of drugs. Lean body mass decreases while fatty tissues increase. The plasma concentrations of water-soluble (low volume of distribution) drugs tend to increase, whereas the plasma concentrations of lipid-soluble (high volume of distribution) drugs tend to decrease. Serum albumin decreases while gamma globulins increase, so total plasma protein concentrations essentially remain the same.


TABLE 4-4 Age-Related Changes in Geriatric and Pediatric Patients































Body Part/Function Affected Geriatric Pediatric
Organ blood flow Decreased Increased
Total body water Decreased Greatly increased
Body fat Increased Decreased
Serum proteins Decreased albumin Decreased albumin
  Increased globulins  
Hepatic metabolism Decreased Greatly decreased

The definition of neonate also varies with species and age, but all the determinants of drug disposition are altered as the foal matures.14 Blood flow to the heart and brain is greater and faster, making the foal more susceptible to drug-induced cardiotoxicity and neurotoxicity. Gastrointestinal absorption tends to be decreased as a result of decreased gastric emptying and decreased intestinal peristalsis. Absorption from intramuscular and subcutaneous sites changes as muscle mass and blood flow change. Neonates have less fat and greater total body water (primarily extracellular fluid) than adults. Therefore low Vd drugs (e.g., gentamicin, ketoprofen) distribute into a larger volume, making it necessary to increase the dose to avoid therapeutic failure. Because of low body fat stores, lipid-soluble drugs will have higher plasma concentrations in foals. For example, moxidectin is much more lipid soluble than ivermectin, so it is more easily overdosed in foals.15 Drug elimination by both hepatic metabolism and renal excretion is limited in neonates, so drug dosing intervals need to be increased for many drugs, such as aminoglycosides and nonsteroidal anti-inflammatory drugs.

Only gold members can continue reading. Log In or Register to continue

Stay updated, free articles. Join our Telegram channel

Jun 8, 2016 | Posted by in EQUINE MEDICINE | Comments Off on Pharmacologic Principles

Full access? Get Clinical Tree

Get Clinical Tree app for offline access