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

Determinants of glucocorticoid dosing

Determinants of glucocorticoid dosing
Literature review current through: Jan 2024.
This topic last updated: Apr 18, 2022.

INTRODUCTION — Natural and synthetic glucocorticoids (also called steroids) can be used for a variety of disorders. These agents are most commonly given in pharmacologic doses to manage conditions that require the suppression of inflammation. Less often, they are used to establish the diagnosis and cause of Cushing's syndrome and for hormone replacement in adrenal insufficiency and congenital adrenal hyperplasia.

The determinants of glucocorticoid dosing include:

Pharmacokinetics of the different drug preparations

Effects of underlying disorders on drug kinetics

Interactions of glucocorticoids with concurrently administered non-glucocorticoid drugs

This topic will review the factors that affect glucocorticoid dosing. A detailed discussion of the structure, absorption, metabolism, and biologic activities of glucocorticoids is presented separately. (See "Pharmacologic use of glucocorticoids".)

BIOEQUIVALENCE AND BIOAVAILABILITY — The different preparations of glucocorticoids are largely bioequivalent (ie, have the same rate of absorption). As an example, most commercially available prednisone tablets are bioequivalent, independent of tablet strength (eg, 1, 2, 5, 10, 20 mg) [1], and the systemic bioavailability of prednisone and prednisolone are similar. Although the rectal and oral absorption of methylprednisolone are variable (relative bioavailability of 50 to 90 percent), such preparations are also bioequivalent [2,3]. Glucocorticoid bioavailability does not appear to be affected, although preeclampsia and gestational hypertension may increase cortisol metabolism (without affecting neonatal birth weight) [4].

Different formulations of glucocorticoid preparations are being developed to improve delivery of these drugs (eg, palmitate large porous particles, oral dissolvable film formulations, nanosuspensions in soft contact lenses, multidose dry powder inhalers), which may improve the usefulness of glucocorticoids [5-8].

Incomplete bioavailability of glucocorticoids has occasionally been noted in certain patients. As an example, in one study 20 percent of patients given methylprednisolone showed poor bioavailability (23 to 65 percent), compared to only 1 of 12 patients given prednisone [9].

The dosing equivalents, relative antiinflammatory and mineralocorticoid activity, and duration of action of the different glucocorticoid preparations, shown in the table (table 1), are discussed separately. (See "Pharmacologic use of glucocorticoids".)

Inhaled glucocorticoids — The bioavailability of inhaled glucocorticoids varies with the physical properties of the particular agent. Eighty percent of inhaled glucocorticoids are swallowed, with the remainder deposited in the lungs. When deposited in the lungs, the more lipophilic compounds (such as fluticasone and beclomethasone) are retained longer in lung tissue.

The absorption of inhaled glucocorticoids also varies with the specific agents. Drugs that are highly lipophilic are relatively poorly absorbed orally (less than 11 percent). By comparison, agents which are not lipophilic (such as budesonide) are somewhat better absorbed orally (less than 20 percent). Since all of the drug deposited in the lung eventually enters the systemic circulation, overall absorption of inhaled glucocorticoids varies between 20 and 40 percent of the administered dose [10].

DISPOSITION — Prednisone is rapidly metabolized via reduction to prednisolone; the elimination of prednisone is approximately 13 times faster than prednisolone [11]. In contrast to cortisol, the endogenously produced glucocorticoid, the synthetic glucocorticoids bind less or minimally to cortisol-binding globulin (CBG, transcortin). Prednisolone has about 60 percent, prednisone 5 percent, and methylprednisolone, dexamethasone, betamethasone, and triamcinolone have less than 1 percent of the affinity of cortisol for CBG. (See "Pharmacologic use of glucocorticoids".)

Clearance — The clearance of prednisolone is 210 mL/min per 1.73 m2, with an elimination half-life of approximately three hours. Clearance decreases with age; as an example, children less than 12 years of age have a 33 percent higher clearance than older children and adults [9].

Glucocorticoids exhibit dose-dependent kinetics. Total prednisolone clearance increases by 75 percent as the intravenous dose increases from 5 to 40 mg [12,13]. Free prednisolone clearances also change with administered dose, but to a lesser degree and require larger doses to demonstrate such kinetics [14]. The clinical consequence of these properties is that a somewhat greater, nonlinear drug effect is observed at prednisolone doses over 40 mg compared to doses between 10 to 20 mg.

Clearances also vary with the time of day. Both prednisolone and methylprednisolone clearance is lower (18 to 28 percent) in the morning than the evening [15,16]. This property, in combination with the disruption of the usual cortisol diurnal rhythm with exogenous glucocorticoids, may result in variations in efficacy when glucocorticoids are administered at different times during the day [17,18]. In one study, for example, the efficacy of prednisolone was assessed in seven asthmatic patients in whom the drug was given at 8 AM and 3 PM, and at 3 PM and 8 PM [19]. The earlier dosing regimen was more effective in improving nocturnal pulmonary function and symptoms.

Distribution in breast milk — Glucocorticoids are excreted in small amounts in human milk. In one study, approximately 0.23 percent of a 5 mg prednisolone dose was found in breast milk [20].

DISEASES AND ALTERED PHYSIOLOGIC STATES — The pharmacokinetics of glucocorticoids vary with certain diseases and pathophysiologic conditions.

Asthma — There are no clinically important pharmacokinetic differences between asthmatic patients and healthy subjects [21].

Cystic fibrosis — The bioavailability of prednisolone appears to be unaffected by cystic fibrosis, but the total prednisolone clearance was increased by over 50 percent [22]. As a result, more frequent dosing may be necessary.

End-stage kidney disease — Among patients treated with hemodialysis, the clearance of total prednisolone is dose-dependent, while the clearance of unbound prednisolone is constant [23]. Hemodialysis also removes significant amounts of methylprednisolone, thereby resulting in a 32 percent reduction in plasma half-life relative to those without ESKD [24]. However, these changes are not sufficient to require a dose adjustment. The removal rate of unbound cortisol in patients treated with peritoneal dialysis is similar to those without ESKD [24,25].

Hyperthyroidism — The clearance of prednisolone is increased in hyperthyroidism. In one small study, total prednisolone clearance was increased by 58 percent and nonrenal clearance (principally hepatic) was increased by 84 percent [26]. There were also small changes in absorption and binding.

Inflammatory bowel disease — The overall kinetics of prednisolone do not appear to be affected by inflammatory bowel disease (IBD). One study, for example, found that total and unbound kinetics of prednisolone were unchanged in active and inactive IBD, although the unbound fraction was increased in active disease [27].

Nephrotic syndrome — Patients have low serum concentrations of albumin and cortisol-binding globulin. However, although their bound and therefore total glucocorticoid concentrations are reduced, the physiologically important unbound (free) serum concentrations of prednisone and prednisolone are similar to nonnephrotic individuals.

Perhaps because of the differences in protein-binding, nonrenal clearance is higher and renal clearance is lower in patients with nephrotic syndrome than in those without nephrotic syndrome [28]. The total prednisolone clearances are higher in nephrotic patients, since the increase in nonrenal clearance is of greater magnitude than the reduction in renal clearance.

Obesity — Obesity can affect the uptake, storage, and metabolism of glucocorticoids, although results are somewhat contradictory.

In one study, the volume of distribution and clearance of prednisone in a person with obesity weighing more than 133 percent of ideal body weight was 20 to 30 percent higher than in a person without obesity [29].

Another report indicated that each 1 percent higher baseline body mass index (BMI) was associated with a 2.9 percent decline in wake-up and total area under curve (AUC) cortisol, suggesting that a higher BMI suppresses cortisol [30].

By contrast, two other reports that examined the metabolism of methylprednisolone and dexamethasone found that clearance among patients with obesity was decreased by about 40 percent when compared with those without obesity [31,32]. One of these studies also examined pharmacodynamic measures of glucocorticoid activity, particularly histamine and T cell responses [31]. Although not statistically significant, there was a clear trend toward an enhanced effect in obese subjects at the same mg/kg dose.

Dosing of glucocorticoids in a person with obesity should be based upon the ideal, rather than total, body weight.

Pregnancy — Little active prednisolone crosses the placenta to the fetus, since the placenta inactivates the drug. However, dexamethasone clearance is increased approximately twofold when compared to nonpregnant women, probably due to enzyme induction [33]. A more detailed discussion of the effects of glucocorticoids on the fetus during pregnancy can be found elsewhere. (See "Safety of rheumatic disease medication use during pregnancy and lactation", section on 'Glucocorticoids'.)

When an effect on the fetus is desired, for example to speed fetal lung maturity in situations where premature delivery is anticipated, fluorinated glucocorticoids are used because these cross the placenta and have pharmacologic effects on the fetal organs. Dexamethasone and betamethasone are often used in this setting. Use of antenatal betamethasone is associated with a lower risk of neonatal death than is dexamethasone [34].

Severe liver disease — In the presence of severe liver disease, the activation of prednisone via metabolism to 6-beta-hydroxyl compounds may be impaired, potentially affecting the efficacy of glucocorticoid therapy [35]. (See "Pharmacologic use of glucocorticoids".)

Patients who undergo liver transplantation because of hepatic C viral infection should receive the lowest dose of glucocorticoids following transplantation that is feasible. The cumulative glucocorticoid dose is correlated closely with post-transplantation hepatitis-C viral load and with mortality rates [36].

DRUG INTERACTIONS

Systemic glucocorticoids — Glucocorticoids undergo metabolism in the liver and other tissues by cytochrome P450 3A4 (CYP 3A4) and other transformations. In vitro data suggest that dexamethasone, methylprednisolone and prednisolone are also substrates of P-glycoprotein membrane efflux transporters. Medications that strongly inhibit or induce CYP 3A4 and/or P-glycoprotein transporters may significantly alter the glucocorticoid serum concentration [37,38].

Medications that increase the systemic glucocorticoid concentration include estrogen derivatives, such as oral contraceptives [39-42] and strong inhibitors of CYP 3A4 (table 2) including some antibiotics (eg, clarithromycin, ritonavir, telaprevir, telithromycin) [43-45], and antifungals (eg, posaconazole, voriconazole) [38,46,47].

Medications that reduce the systemic glucocorticoid concentration include aluminum/magnesium containing antacids, which decrease prednisone bioavailability due to decreased oral absorption [48,49], and strong inducers of CYP 3A4 (eg, carbamazepine, phenobarbital, phenytoin and rifampin) [50-55].

However, a number of agents often used with glucocorticoids appear to have no substantial interaction with them. These include azathioprine, methotrexate, histamine2 antagonists (eg, famotidine, cimetidine, ranitidine), proton pump inhibitors (eg, omeprazole, pantoprazole, rabeprazole), and diazepam [50,56-60].

The major drug interactions with systemic glucocorticoids, a summary of effect(s), and management suggestions are listed in the table (table 3). For additional interactions, see the Lexicomp drug interactions program included with UpToDate.

Inhaled or intranasal glucocorticoids — Inhibitors of CYP 3A4 may impair the metabolism of glucocorticoids administered intranasally or by inhalation and may increase the serum concentration associated with these methods of administration. Use caution when combining high-dose intranasal or inhaled glucocorticoids with strong CYP 3A4 inhibitors (table 2). This is discussed elsewhere. (See "Pharmacotherapy of allergic rhinitis" and "Major side effects of inhaled glucocorticoids", section on 'Medication interactions'.)

SOCIETY GUIDELINE LINKS — Links to society and government-sponsored guidelines from selected countries and regions around the world are provided separately. (See "Society guideline links: Side effects of anti-inflammatory and anti-rheumatic drugs".)

SUMMARY AND RECOMMENDATIONS

The determinants of glucocorticoid dosing include pharmacokinetics of the different drug preparations, effects of underlying disorders on drug kinetics, and interactions of glucocorticoids with concurrently administered non-glucocorticoid drugs. (See 'Introduction' above.)

The different preparations of glucocorticoids are largely bioequivalent (ie, have the same rate of absorption). The systemic bioavailability of prednisone and prednisolone are similar; the rectal and oral absorption of methylprednisolone are variable (relative bioavailability of 50 to 90 percent). Incomplete bioavailability of glucocorticoids has occasionally been noted. The bioavailability of inhaled glucocorticoids varies with the physical properties of the particular agent. (See 'Bioequivalence and bioavailability' above and 'Inhaled glucocorticoids' above.)

The dosing equivalents, relative antiinflammatory and mineralocorticoid activity, and duration of action of the different glucocorticoid preparations, shown in the table (table 1), are discussed separately. (See "Pharmacologic use of glucocorticoids".)

Prednisone is rapidly metabolized via reduction to prednisolone. In contrast to cortisol, the endogenously-produced glucocorticoid, the synthetic glucocorticoids bind less or minimally to cortisol-binding globulin. Prednisolone has an elimination half-life of approximately three hours; clearance decreases with age and varies with time of day. Kinetics are dose-dependent. Only small amounts are excreted in breast milk. (See 'Disposition' above and 'Clearance' above and 'Distribution in breast milk' above.)

The pharmacokinetics of glucocorticoids vary with certain diseases and pathophysiologic conditions, including cystic fibrosis, end-stage kidney disease and hemodialysis, hyperthyroidism, nephrotic syndrome, obesity, pregnancy, and severe liver disease. (See 'Diseases and altered physiologic states' above.)

Significant interactions with glucocorticoids have been documented for the medications listed in the table (table 3); a change of less than 30 percent is not likely to be clinically significant. Agents often used with glucocorticoids that appear to have no substantial interaction with them include azathioprine, cyclosporine, methotrexate, cimetidine, ranitidine, theophylline, and diazepam. (See 'Drug interactions' above.)

Drugs that reduce the systemic glucocorticoid concentration include large doses of aluminum/magnesium hydroxide, which decrease bioavailability, and most anticonvulsants, which enhance glucocorticoid metabolism. Drugs that raise the systemic glucocorticoid concentration include some oral contraceptives, and certain antibiotics and antifungal agents; these drugs decrease metabolizing enzymes. (See 'Drug interactions' above.)

  1. Francisco GE, Honigberg IL, Stewart JT, et al. In vitro and in vivo bioequivalence of commercial prednisone tablets. Biopharm Drug Dispos 1984; 5:335.
  2. Lima JJ, Giller J, Mackichan JJ, Jusko WJ. Bioavailability of hydrocortisone retention enemas in normal subjects. Am J Gastroenterol 1980; 73:232.
  3. Garg DC, Wagner JG, Sakmar E, et al. Rectal and oral absorption of methylprednisolone acetate. Clin Pharmacol Ther 1979; 26:232.
  4. Kosicka K, Siemiątkowska A, Krzyścin M, et al. Glucocorticoid Metabolism in Hypertensive Disorders of Pregnancy: Analysis of Plasma and Urinary Cortisol and Cortisone. PLoS One 2015; 10:e0144343.
  5. N'Guessan A, Fattal E, Chapron D, et al. Dexamethasone palmitate large porous particles: A controlled release formulation for lung delivery of corticosteroids. Eur J Pharm Sci 2018; 113:185.
  6. Diamant Z, Samuelsson Palmgren G, Westrin B, Bjermer L. Phase I study evaluating the safety, tolerability and pharmacokinetics of a novel oral dissolvable film containing dexamethasone versus Fortecortin dexamethasone tablets. Eur Clin Respir J 2017; 4:1353395.
  7. García-Millán E, Quintáns-Carballo M, Otero-Espinar FJ. Improved release of triamcinolone acetonide from medicated soft contact lenses loaded with drug nanosuspensions. Int J Pharm 2017; 525:226.
  8. Bernstein DI, Gillespie M, Song S, Steinfeld J. Safety, efficacy, and dose response of fluticasone propionate delivered via the novel MDPI in patients with severe asthma: A randomized, controlled, dose-ranging study. J Asthma 2017; 54:559.
  9. Hill MR, Szefler SJ, Ball BD, et al. Monitoring glucocorticoid therapy: a pharmacokinetic approach. Clin Pharmacol Ther 1990; 48:390.
  10. Johnson M. Pharmacodynamics and pharmacokinetics of inhaled glucocorticoids. J Allergy Clin Immunol 1996; 97:169.
  11. Hale VG, Aizawa K, Sheiner LB, Benet LZ. Disposition of prednisone and prednisolone in the perfused rabbit liver: modeling hepatic metabolic processes. J Pharmacokinet Biopharm 1991; 19:597.
  12. Rose JQ, Yurchak AM, Jusko WJ. Dose dependent pharmacokinetics of prednisone and prednisolone in man. J Pharmacokinet Biopharm 1981; 9:389.
  13. Legler UF, Frey FJ, Benet LZ. Prednisolone clearance at steady state in man. J Clin Endocrinol Metab 1982; 55:762.
  14. Wald JA, Law RM, Ludwig EA, et al. Evaluation of dose-related pharmacokinetics and pharmacodynamics of prednisolone in man. J Pharmacokinet Biopharm 1992; 20:567.
  15. Meffin PJ, Wing LM, Sallustio BC, Brooks PM. Alterations in prednisolone disposition as a result of oral contraceptive use and dose. Br J Clin Pharmacol 1984; 17:655.
  16. Fisher LE, Ludwig EA, Wald JA, et al. Pharmacokinetics and pharmacodynamics of methylprednisolone when administered at 8 am versus 4 pm. Clin Pharmacol Ther 1992; 51:677.
  17. Sugita ET, Niebergall PJ. Bioavailability monograph: Prednisolone. J Am Pharm Assoc 1975; NS15:529.
  18. Theissen JJ. Bioavailability monograph: Prednisolone. J Am Pharm Assoc 1976; NS16:143.
  19. Reinberg A, Gervais P, Chaussade M, et al. Circadian changes in effectiveness of corticosteroids in eight patients with allergic asthma. J Allergy Clin Immunol 1983; 71:425.
  20. Brooks PM, Needs CJ. The use of antirheumatic medication during pregnancy and in the puerperium. Rheum Dis Clin North Am 1989; 15:789.
  21. Mortimer O, Grettve L, Lindström B, et al. Bioavailability of prednisolone in asthmatic patients with a poor response to steroid treatment. Eur J Respir Dis 1987; 71:372.
  22. Dove AM, Szefler SJ, Hill MR, et al. Altered prednisolone pharmacokinetics in patients with cystic fibrosis. J Pediatr 1992; 120:789.
  23. Frey FJ, Gambertoglio JG, Frey BM, et al. Nonlinear plasma protein binding and haemodialysis clearance of prednisolone. Eur J Clin Pharmacol 1982; 23:65.
  24. Sherlock JE, Letteri JM. Effect of hemodialysis on methylprednisolone plasma levels. Nephron 1977; 18:208.
  25. Zager PG, Spalding CT, Frey HJ, et al. Dialysance of adrenocorticoids during continuous ambulatory peritoneal dialysis. J Clin Endocrinol Metab 1988; 67:110.
  26. Frey FJ, Horber FF, Frey BM. Altered metabolism and decreased efficacy of prednisolone and prednisone in patients with hyperthyroidism. Clin Pharmacol Ther 1988; 44:510.
  27. Milsap RL, George DE, Szefler SJ, et al. Effect of inflammatory bowel disease on absorption and disposition of prednisolone. Dig Dis Sci 1983; 28:161.
  28. Frey FJ, Frey BM. Altered plasma protein-binding of prednisolone in patients with the nephrotic syndrome. Am J Kidney Dis 1984; 3:339.
  29. Milsap RL, Plaisance KI, Jusko WJ. Prednisolone disposition in obese men. Clin Pharmacol Ther 1984; 36:824.
  30. Joseph JJ, Wang X, Diez Roux AV, et al. Antecedent longitudinal changes in body mass index are associated with diurnal cortisol curve features: The multi-ethnic study of atherosclerosis. Metabolism 2017; 68:95.
  31. Dunn TE, Ludwig EA, Slaughter RL, et al. Pharmacokinetics and pharmacodynamics of methylprednisolone in obesity. Clin Pharmacol Ther 1991; 49:536.
  32. Lamiable D, Vistelle R, Sulmont V, et al. [Pharmacokinetics of dexamethasone administered orally in obese patients]. Therapie 1990; 45:311.
  33. Tsuei SE, Petersen MC, Ashley JJ, et al. Disporition of synthetic glucocorticoids. II. Dexamethasone in parturient women. Clin Pharmacol Ther 1980; 28:88.
  34. Jobe AH, Soll RF. Choice and dose of corticosteroid for antenatal treatments. Am J Obstet Gynecol 2004; 190:878.
  35. Renner E, Horber FF, Jost G, et al. Effect of liver function on the metabolism of prednisone and prednisolone in humans. Gastroenterology 1986; 90:819.
  36. Charlton M. Management of recurrence of hepatitis C infection following liver transplantation. Minerva Chir 2003; 58:717.
  37. Gary H. Wynn. Transplant surgery and rheumatology. In: Clinical manual of drug interaction principles for medical practice, 1st ed, Gary H. Wynn (Ed), APA Publishing Inc., Washington DC 2009. p.462.
  38. Czock D, Keller F, Rasche FM, Häussler U. Pharmacokinetics and pharmacodynamics of systemically administered glucocorticoids. Clin Pharmacokinet 2005; 44:61.
  39. Olivesi A. Modified elimination of prednisolone in epileptic patients on carbamazepine monotherapy, and in women using low-dose oral contraceptives. Biomed Pharmacother 1986; 40:301.
  40. Boekenoogen SJ, Szefler SJ, Jusko WJ. Prednisolone disposition and protein binding in oral contraceptive users. J Clin Endocrinol Metab 1983; 56:702.
  41. Gustavson LE, Legler UF, Benet LZ. Impairment of prednisolone disposition in women taking oral contraceptives or conjugated estrogens. J Clin Endocrinol Metab 1986; 62:234.
  42. Frey BM, Schaad HJ, Frey FJ. Pharmacokinetic interaction of contraceptive steroids with prednisone and prednisolone. Eur J Clin Pharmacol 1984; 26:505.
  43. LaForce CF, Szefler SJ, Miller MF, et al. Inhibition of methylprednisolone elimination in the presence of erythromycin therapy. J Allergy Clin Immunol 1983; 72:34.
  44. Szefler SJ, Brenner M, Jusko WJ, et al. Dose- and time-related effect of troleandomycin on methylprednisolone elimination. Clin Pharmacol Ther 1982; 32:166.
  45. Szefler SJ, Ellis EF, Brenner M, et al. Steroid-specific and anticonvulsant interaction aspects of troleandomycin-steroid therapy. J Allergy Clin Immunol 1982; 69:455.
  46. Zürcher RM, Frey BM, Frey FJ. Impact of ketoconazole on the metabolism of prednisolone. Clin Pharmacol Ther 1989; 45:366.
  47. Yamashita SK, Ludwig EA, Middleton E Jr, Jusko WJ. Lack of pharmacokinetic and pharmacodynamic interactions between ketoconazole and prednisolone. Clin Pharmacol Ther 1991; 49:558.
  48. Tanner AR, Caffin JA, Halliday JW, Powell LW. Concurrent administration of antacids and prednisone: effect on serum levels of prednisolone. Br J Clin Pharmacol 1979; 7:397.
  49. Uribe M, Casian C, Rojas S, et al. Decreased bioavailability of prednisone due to antacids in patients with chronic active liver disease and in healthy volunteers. Gastroenterology 1981; 80:661.
  50. Stjernholm MR, Katz FH. Effects of diphenylhydantoin, phenobarbital, and diazepam on the metabolism of methylprednisolone and its sodium succinate. J Clin Endocrinol Metab 1975; 41:887.
  51. Frey FJ, Frey BM. Urinary 6 beta-hydroxyprednisolone excretion indicates enhanced prednisolone catabolism. J Lab Clin Med 1983; 101:593.
  52. Brooks SM, Werk EE, Ackerman SJ, et al. Adverse effects of phenobarbital on corticosteroid metabolism in patients with bronchial asthma. N Engl J Med 1972; 286:1125.
  53. Frey BM, Frey FJ. Phenytoin modulates the pharmacokinetics of prednisolone and the pharmacodynamics of prednisolone as assessed by the inhibition of the mixed lymphocyte reaction in humans. Eur J Clin Invest 1984; 14:1.
  54. Evans PJ, Walker RF, Peters JR, et al. Anticonvulsant therapy and cortisol elimination. Br J Clin Pharmacol 1985; 20:129.
  55. Petereit LB, Meikle AW. Effectiveness of prednisolone during phenytoin therapy. Clin Pharmacol Ther 1977; 22:912.
  56. Frey FJ, Lozada F, Guentert T, Frey BM. A single dose of azathioprine does not affect the pharmacokinetics of prednisolone following oral prednisone. Eur J Clin Pharmacol 1981; 19:209.
  57. Frey FJ, Schnetzer A, Horber FF, Frey BM. Evidence that cyclosporine does not affect the metabolism of prednisolone after renal transplantation. Transplantation 1987; 43:494.
  58. Glynn-Barnhart AM, Erzurum SC, Leff JA, et al. Effect of low-dose methotrexate on the disposition of glucocorticoids and theophylline. J Allergy Clin Immunol 1991; 88:180.
  59. Sirgo MA, Rocci ML Jr, Ferguson RK, et al. Effects of cimetidine and ranitidine on the conversion of prednisone to prednisolone. Clin Pharmacol Ther 1985; 37:534.
  60. Brooks SM, Sholiton LJ, Werk EE Jr, Altenau P. The effects of ephedrine and theophylline on dexamethasone metabolism in bronchial asthma. J Clin Pharmacol 1977; 17:308.
Topic 7976 Version 20.0

References

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