ﺑﺎﺯﮔﺸﺖ ﺑﻪ ﺻﻔﺤﻪ ﻗﺒﻠﯽ
خرید پکیج
تعداد آیتم قابل مشاهده باقیمانده : 3 مورد
نسخه الکترونیک
medimedia.ir

Pyrazinamide: An overview

Pyrazinamide: An overview
Literature review current through: Jan 2024.
This topic last updated: Dec 13, 2023.

INTRODUCTION — Pyrazinamide (PZA) is an antimicrobial agent that is most commonly used for treatment of active tuberculosis (TB) during the initial phase of therapy (generally the first two months of treatment), in combination with other agents.

The spectrum of PZA is relatively narrow; it demonstrates clinically significant antibacterial activity only against Mycobacterium tuberculosis and Mycobacterium africanum [1].

Issues related to the clinical use of PZA will be reviewed here. Issues related to treatment of latent and active TB are discussed separately. (See related topics.)

MECHANISM OF ACTION — The mechanism of action for PZA is unknown [2,3]. The parent compound enters the bacterium passively and is metabolized via pyrazinamidase (PZase) within the cytoplasm to pyrazinoic acid; pyrazinoic acid is the active form of the drug [4]. PZA and its analog, 5-chloro-PZA, may inhibit the fatty acid synthetase I enzyme of M. tuberculosis [5,6]. PZA is generally considered to be a bacteriostatic agent.

PZA is thought to be more active at an acidic pH (eg, within macrophages). It demonstrates activity against both replicating and slow-growing populations [3]. The role of PZA against intracellular organisms remains uncertain [7]. In one study, pyrazinoic acid remained outside of M. tuberculosis cells at a neutral or alkaline pH but accumulated within cells at an acidic pH [8]. In the same study, Mycobacterium smegmatis (which is not susceptible to PZA) was found to convert PZA to pyrazinoic acid but, due to an active efflux mechanism, did not accumulate the metabolite, even at an acidic pH. Other mycobacterial strains appear to have natural resistance to PZA due to lack of PZase activity or absence of transport mechanisms to take up the drug [4].

When used as part of combination therapy, PZA appears to accelerate the sterilizing effect of isoniazid, rifampin, and bedaquiline [1,9]. In selected populations, this has enabled reduction in the duration of treatment infection due to susceptible M. tuberculosis isolates.

RESISTANCE — PZA is not used as monotherapy in the treatment of M. tuberculosis because of the rapid development of resistance to the drug. The major mechanism of resistance to PZA arises from mutations in the gene encoding PZase, pncA [10]. A number of different mutations in this gene have been demonstrated [11-13]. Rates of tuberculosis resistance to PZA have been reported to be approximately 16 percent [14] but can range from 10 to 85 percent depending on geography and patient population [14].

Phenotypic susceptibility testing for PZA is challenging and (as a result) its role in routine care is not routinely recommended [15,16]. The role of genotypic testing was evaluated in a study including 157 patients with MDR-TB, genotypic PZA resistance was associated with a longer time to sputum culture conversion [17].

PHARMACOKINETICS — Absorption of PZA is thought to be rapid after oral administration, with peak serum concentrations achieved within one to two hours [1,18]. Only minor changes in bioavailability and peak serum concentrations are demonstrated when PZA is taken with meals [18,19]. Serum concentrations have been reported to range from 30 to 50 mcg/mL after a 25 mg/kg dose [1]. Limited data are available to determine PZA outside the blood; modeling with standard exposures report that extrapulmonary concentrations may be below critical concentrations at most sites of infection [20].

Protein binding of PZA is very low (median of 1 percent in one report) [21]. PZA is thought to be widely distributed in many body tissues and fluids, with a volume of distribution estimated to be 0.6 to 0.7 L/kg. Intrapulmonary concentrations of PZA are high (4 to 40 times the reported minimum inhibitory concentration for PZA-susceptible strains in one study) [22]. PZA distributes well into the cerebrospinal fluid, with concentrations reported to be up to 90 percent to those observed in simultaneous serum samples [23]. Intracellular PZA concentrations are similar to those of extracellular fluids. Children appear to have a larger volume of distribution and a slower absorption and clearance of PZA than adults [24].

PZA undergoes extensive hepatic metabolism. Approximately 70 percent of the dose is excreted in the urine in the form of metabolites. The half-life is approximately 9 to 10 hours in patients with normal renal and hepatic function [1]. The drug and its metabolites are removed by hemodialysis [25].

Significant variability of PZA exposure has been reported when dosing is based on body weight. Factors thought to contribute to such variability may include HIV coinfection, age, gender, weight, and organ dysfunction (hepatic or renal).

PHARMACODYNAMICS — Data reporting relationships between PZA concentrations and clinical outcomes are sparse. Studies suggest a relationship between higher maximum concentration and both reduced time to culture conversion and a higher probability of two-month culture conversion [26]. In another report, relationships between PZA area under the time-concentration curve ≤363 mg·hour/L was associated with poor treatment outcome [27].

DOSING AND ADMINISTRATION — Intermittent dosing is performed as part of directly observed therapy (DOT) (table 1).

One study including 72 adults receiving PZA suggested that, based on pharmacodynamic targets (which define the optimal relationship between drug exposure and pathogen susceptibility), the weight-based dosing recommendations described above may be inadequate [28]. This study suggested that optimal dosing may be between 35 and 50mg/kg/day for most patients. However, thus far, such dosing has not been adopted in published guidelines. Higher doses of PZA (25 to 40 mg/kg daily) have also been proposed in patients with multidrug-resistant strains [29].

SPECIAL POPULATIONS

Renal insufficiency — PZA dosing requires adjustment in patients with severe renal insufficiency. In patients with a creatinine clearance <30 mL/min or in patients with end-stage kidney disease requiring hemodialysis, 25 to 35 mg/kg per dose should be administered three times a week (after hemodialysis if applicable). Since PZA and pyrazinoic acid are removed by hemodialysis, doses should be administered after dialysis [30]. PZA increases the risk of hyperuricemia in patients with renal insufficiency.

Liver dysfunction — Dose adjustment may be required in patients with hepatic dysfunction, though there are no published guidelines available for administration of PZA to patients with significant liver disease.

Children — Issues related to PZA dosing in children are discussed separately. (See "Tuberculosis disease in children: Treatment and prevention".)

Pregnancy — Issues related to use of PZA in pregnancy are discussed separately. (See "Tuberculosis disease (active tuberculosis) in pregnancy".)

Breastfeeding — Because of the low concentrations of PZA expected in breastmilk, the United States Centers for Disease Control and Prevention does not discourage breastfeeding in women taking PZA [31].

Obesity — Weight-based dosing of PZA is based on estimation of lean body weight. Dosing recommendations for obese patients are not available [29].

HIV coinfection — Pharmacokinetic studies of PZA in patients coinfected with HIV have reported subtherapeutic PZA, which was impacted by concomitant use of antiretroviral therapy. The authors recommended PZA dosing be based on gender and serum creatinine in this patient population [32].

ADVERSE EFFECTS — The most clinically significant adverse events associated with PZA administration are hepatotoxicity, gastrointestinal intolerance, nongouty polyarthralgias, and asymptomatic hyperuricemia.

Hepatotoxicity — Hepatotoxicity due to PZA can manifest as elevated aspartate aminotransferase concentration, jaundice, and liver tenderness. Hepatotoxicity appears to be dose and duration-related, and is infrequent at daily doses of ≤25 mg/kg per day when administered in multi-drug regimens for active tuberculosis (TB) [26]. Prior liver disease, age over 60 years, and concomitant use of isoniazid and rifampin appear to increase risk for development of hepatotoxicity [33,34].

Unexpectedly high rates of hepatotoxicity have been observed with administration of PZA in combination with rifampin for treatment of latent TB infection (LTBI). The mechanism for this enhanced toxicity is not clear; revised guidelines from the United States Centers for Disease Control recommend against the use of this combination for treating LTBI [30,35]. Some reports have suggested that PZA used with a second drug for treating LTBI may also potentiate hepatotoxicity.

Other reactions — Pyrazinoic acid, the active metabolite of PZA, competes with uric acid for elimination via the kidneys, resulting in asymptomatic hyperuricemia and exacerbations of gout in susceptible individuals [36]. Labile blood glucose concentrations have been observed in diabetic patients. Gastrointestinal intolerance may result in nausea and vomiting. Dermatologic (maculopapular rash, urticaria, photosensitivity) and hematologic (thrombocytopenia, sideroblastic anemia) side effects have been described but are relatively rare. (See "Sideroblastic anemias: Diagnosis and management", section on 'Medications'.)

MONITORING — Patients receiving PZA (in combination with other antituberculous agents) should undergo baseline assessment and periodic monitoring of liver function. Serum uric acid measurements may be obtained in individuals who have a history of gout, who are receiving concomitant medications that alter uric acid excretion (such as loop diuretics), or who are at increased risk for hyperuricemia due to underlying renal dysfunction.

PZA serum concentration monitoring is not routinely performed but may be considered in patients with delayed or poor response to TB treatment or relapse after therapy is complete. Those with gastrointestinal disorders, diabetes mellitus, HIV co-infection, or malnutrition may experience reductions in drug absorption. In addition, those patients with risk factors for PZA-related toxicity may also be considered. While the data to support a target concentration are sparse, a proposed peak therapeutic serum concentration range two hours following administration (Cmax) is 20 to 60 mcg/mL [37].

DRUG INTERACTIONS — Combination therapy with rifampin and PZA has been associated with severe and fatal hepatotoxic reactions.

SUMMARY

Pyrazinamide (PZA) is an antimicrobial agent used for treatment of active tuberculosis (TB) during the initial phase of therapy (generally the first two months of treatment), in combination with other agents. PZA is not used as monotherapy in the treatment of Mycobacterium tuberculosis because of rapid development of resistance to the drug. (See 'Introduction' above.)

The mechanism of action for PZA is not fully understood; it may inhibit the fatty acid synthetase I enzyme of M. tuberculosis. PZA is thought to be more active in acidic environments (eg, within macrophages). When administered as part of combination therapy, PZA appears to accelerate the sterilizing effect of isoniazid and rifampin. (See 'Mechanism of action' above.)

PZA is thought to be widely distributed in many body tissues and fluids. Intrapulmonary concentrations of PZA are high. PZA also distributes well into the cerebrospinal fluid. PZA undergoes extensive hepatic metabolism; approximately 70 percent of the dose is excreted in the urine in the form of metabolites. The half-life is approximately 9 to 10 hours in patients with normal renal and hepatic function. (See 'Pharmacokinetics' above.)

Dosing is summarized in the table (table 1). PZA dosing requires adjustment in patients with severe renal insufficiency and may also be required in patients with hepatic dysfunction. For patients on hemodialysis, the PZA should be administered after dialysis. (See 'Dosing and administration' above.)

Adverse effects associated with PZA administration include hepatotoxicity, gastrointestinal intolerance, polyarthralgia, and asymptomatic hyperuricemia. PZA should not be used with rifampin for treating latent TB infection due to high rates of hepatotoxicity. PZA increases the risk of hyperuricemia most commonly in the setting of renal insufficiency or with the use of concomitant medications, which decrease uric acid excretion. (See 'Adverse effects' above.)

Patients receiving PZA should undergo baseline assessment and periodic monitoring of liver function. Serum uric acid measurements may be obtained in individuals who have a history of gout, who are receiving concomitant medications that decrease uric acid excretion (such as loop diuretics), or who are at increased risk for hyperuricemia due to underlying renal dysfunction. (See 'Monitoring' above.)

  1. Heifets L. Antimycobacterial agents: Pyrazinamide. In: Antimicrobial Therapy and Vaccines, Yu VL, Merigan TC, Barriere SL (Eds), Williams and Wilkins, 1999.
  2. Anthony RM, den Hertog AL, van Soolingen D. 'Happy the man, who, studying nature's laws, Thro' known effects can trace the secret cause.' Do we have enough pieces to solve the pyrazinamide puzzle? J Antimicrob Chemother 2018; 73:1750.
  3. Lamont EA, Dillon NA, Baughn AD. The Bewildering Antitubercular Action of Pyrazinamide. Microbiol Mol Biol Rev 2020; 84.
  4. Raynaud C, Lanéelle MA, Senaratne RH, et al. Mechanisms of pyrazinamide resistance in mycobacteria: importance of lack of uptake in addition to lack of pyrazinamidase activity. Microbiology 1999; 145 ( Pt 6):1359.
  5. Zimhony O, Cox JS, Welch JT, et al. Pyrazinamide inhibits the eukaryotic-like fatty acid synthetase I (FASI) of Mycobacterium tuberculosis. Nat Med 2000; 6:1043.
  6. Ngo SC, Zimhony O, Chung WJ, et al. Inhibition of isolated Mycobacterium tuberculosis fatty acid synthase I by pyrazinamide analogs. Antimicrob Agents Chemother 2007; 51:2430.
  7. Heifets L, Higgins M, Simon B. Pyrazinamide is not active against Mycobacterium tuberculosis residing in cultured human monocyte-derived macrophages. Int J Tuberc Lung Dis 2000; 4:491.
  8. Zhang Y, Scorpio A, Nikaido H, Sun Z. Role of acid pH and deficient efflux of pyrazinoic acid in unique susceptibility of Mycobacterium tuberculosis to pyrazinamide. J Bacteriol 1999; 181:2044.
  9. Muliaditan M, Della Pasqua O. Evaluation of pharmacokinetic-pharmacodynamic relationships and selection of drug combinations for tuberculosis. Br J Clin Pharmacol 2021; 87:140.
  10. Ramaswamy S, Musser JM. Molecular genetic basis of antimicrobial agent resistance in Mycobacterium tuberculosis: 1998 update. Tuber Lung Dis 1998; 79:3.
  11. Sachais BS, Nachamkindagger I I, Mills JK, Leonard DG. Novel pncA Mutations in Pyrazinamide-Resistant Isolates of Mycobacterium tuberculosis. Mol Diagn 1998; 3:229.
  12. Lemaitre N, Sougakoff W, Truffot-Pernot C, Jarlier V. Characterization of new mutations in pyrazinamide-resistant strains of Mycobacterium tuberculosis and identification of conserved regions important for the catalytic activity of the pyrazinamidase PncA. Antimicrob Agents Chemother 1999; 43:1761.
  13. Marttila HJ, Marjamäki M, Vyshnevskaya E, et al. pncA mutations in pyrazinamide-resistant Mycobacterium tuberculosis isolates from northwestern Russia. Antimicrob Agents Chemother 1999; 43:1764.
  14. Whitfield MG, Soeters HM, Warren RM, et al. A Global Perspective on Pyrazinamide Resistance: Systematic Review and Meta-Analysis. PLoS One 2015; 10:e0133869.
  15. Mok S, Roycroft E, Flanagan PR, et al. Overcoming the Challenges of Pyrazinamide Susceptibility Testing in Clinical Mycobacterium tuberculosis Isolates. Antimicrob Agents Chemother 2021; 65:e0261720.
  16. Domínguez J, Boeree MJ, Cambau E, et al. Clinical implications of molecular drug resistance testing for Mycobacterium tuberculosis: a 2023 TBnet/RESIST-TB consensus statement. Lancet Infect Dis 2023; 23:e122.
  17. Kuhlin J, Davies Forsman L, Mansjö M, et al. Genotypic Resistance of Pyrazinamide but Not Minimum Inhibitory Concentration Is Associated With Longer Time to Sputum Culture Conversion in Patients With Multidrug-resistant Tuberculosis. Clin Infect Dis 2021; 73:e3511.
  18. Peloquin CA, Bulpitt AE, Jaresko GS, et al. Pharmacokinetics of pyrazinamide under fasting conditions, with food, and with antacids. Pharmacotherapy 1998; 18:1205.
  19. Saktiawati AM, Sturkenboom MG, Stienstra Y, et al. Impact of food on the pharmacokinetics of first-line anti-TB drugs in treatment-naive TB patients: a randomized cross-over trial. J Antimicrob Chemother 2016; 71:703.
  20. Ramachandran A, Gadgil CJ. A physiologically-based pharmacokinetic model for tuberculosis drug disposition at extrapulmonary sites. CPT Pharmacometrics Syst Pharmacol 2023; 12:1274.
  21. Alghamdi WA, Al-Shaer MH, Peloquin CA. Protein Binding of First-Line Antituberculosis Drugs. Antimicrob Agents Chemother 2018; 62.
  22. Conte JE Jr, Golden JA, Duncan S, et al. Intrapulmonary concentrations of pyrazinamide. Antimicrob Agents Chemother 1999; 43:1329.
  23. Stemkens R, Litjens CHC, Dian S, et al. Pharmacokinetics of pyrazinamide during the initial phase of tuberculous meningitis treatment. Int J Antimicrob Agents 2019; 54:371.
  24. Roy V, Tekur U, Chopra K. Pharmacokinetics of pyrazinamide in children suffering from pulmonary tuberculosis. Int J Tuberc Lung Dis 1999; 3:133.
  25. Malone RS, Fish DN, Spiegel DM, et al. The effect of hemodialysis on isoniazid, rifampin, pyrazinamide, and ethambutol. Am J Respir Crit Care Med 1999; 159:1580.
  26. Zhang N, Savic RM, Boeree MJ, et al. Optimising pyrazinamide for the treatment of tuberculosis. Eur Respir J 2021; 58.
  27. Pasipanodya JG, McIlleron H, Burger A, et al. Serum drug concentrations predictive of pulmonary tuberculosis outcomes. J Infect Dis 2013; 208:1464.
  28. Alsultan A, Savic R, Dooley KE, et al. Population Pharmacokinetics of Pyrazinamide in Patients with Tuberculosis. Antimicrob Agents Chemother 2017; 61.
  29. Nahid P, Mase SR, Migliori GB, et al. Treatment of Drug-Resistant Tuberculosis. An Official ATS/CDC/ERS/IDSA Clinical Practice Guideline. Am J Respir Crit Care Med 2019; 200:e93.
  30. Centers for Disease Control and Prevention (CDC), American Thoracic Society. Update: adverse event data and revised American Thoracic Society/CDC recommendations against the use of rifampin and pyrazinamide for treatment of latent tuberculosis infection--United States, 2003. MMWR Morb Mortal Wkly Rep 2003; 52:735.
  31. Drugs and Lactation Database (LactMed), National Library of Medicine (US).
  32. Sundell J, Wijk M, Bienvenu E, et al. Factors Affecting the Pharmacokinetics of Pyrazinamide and Its Metabolites in Patients Coinfected with HIV and Implications for Individualized Dosing. Antimicrob Agents Chemother 2021; 65:e0004621.
  33. Schaberg T, Rebhan K, Lode H. Risk factors for side-effects of isoniazid, rifampin and pyrazinamide in patients hospitalized for pulmonary tuberculosis. Eur Respir J 1996; 9:2026.
  34. Chang KC, Leung CC, Yew WW, et al. Hepatotoxicity of pyrazinamide: cohort and case-control analyses. Am J Respir Crit Care Med 2008; 177:1391.
  35. Jasmer RM, Saukkonen JJ, Blumberg HM, et al. Short-course rifampin and pyrazinamide compared with isoniazid for latent tuberculosis infection: a multicenter clinical trial. Ann Intern Med 2002; 137:640.
  36. Koumbaniou C, Nicopoulos C, Vassiliou M, et al. Is pyrazinamide really the third drug of choice in the treatment of tuberculosis? Int J Tuberc Lung Dis 1998; 2:675.
  37. Alsultan A, Peloquin CA. Therapeutic drug monitoring in the treatment of tuberculosis: an update. Drugs 2014; 74:839.
Topic 492 Version 26.0

References

آیا می خواهید مدیلیب را به صفحه اصلی خود اضافه کنید؟