farmacocinética corticoide.pdf - UFF

farmacocinética corticoide.pdf - UFF farmacocinética corticoide.pdf - UFF

23.02.2014 Views

Clin Pharmacokinet 2005; 44 (1): 61-98 REVIEW ARTICLE 0312-5963/05/0001-0061/$34.95/0 © 2005 Adis Data Information BV. All rights reserved. Pharmacokinetics and Pharmacodynamics of Systemically Administered Glucocorticoids David Czock, Frieder Keller, Franz Maximilian Rasche and Ulla Häussler Division of Nephrology, University Hospital Ulm, Ulm, Germany Contents Abstract .....................................................................................62 1. Pharmacokinetics .........................................................................63 1.1 General ..............................................................................63 1.1.1 Absorption and Distribution ......................................................63 1.1.2 Metabolism and Excretion .......................................................66 1.2 Selected Examples ....................................................................66 1.2.1 Cortisol, Cortisone, and Hydrocortisone ...........................................66 1.2.2 Prednisolone and Prednisone .....................................................66 1.2.3 Methylprednisolone .............................................................67 1.2.4 Dexamethasone ................................................................67 1.3 Pharmacokinetic Drug-Drug Interactions ................................................68 1.3.1 Influence of Other Drugs on Glucocorticoid Pharmacokinetics ......................68 1.3.2 Influence of Glucocorticoids on the Pharmacokinetics of Other Drugs ................69 1.3.3 Interactions Between Glucocorticoids and Ciclosporin, Tacrolimus or Sirolimus (Rapamycin) ...................................................................69 1.4 Influence of Diseases on Glucocorticoid Pharmacokinetics ...............................70 2. Genomic and Nongenomic Mechanisms of Glucocorticoid Effects ............................71 2.1 Genomic Mechanisms ................................................................71 2.1.1 Glucocorticoid Receptors ........................................................71 2.1.2 Glucocorticoid Response Elements ...............................................71 2.1.3 Transcription Factors (Nuclear Factor κB and Activator Protein 1) ....................73 2.1.4 Post-Transcriptional and Translational Mechanisms ..................................73 2.2 Nongenomic Mechanisms .............................................................73 2.2.1 Specific Nongenomic Mechanisms ...............................................74 2.2.2 Nonspecific Nongenomic Mechanisms ............................................75 3. The Host Defence Response and the Effects of Glucocorticoids ...............................75 3.1 Molecular Mechanisms ................................................................75 3.1.1 Cytokines and Chemokines ......................................................76 3.1.2 Inflammatory Enzymes ...........................................................76 3.2 Cellular Mechanisms ..................................................................76 3.2.1 Cell Trafficking and Adhesion Molecules ...........................................76 3.2.2 T Cell Differentiation .............................................................76 3.2.3 Cell Proliferation ................................................................77 3.2.4 Apoptosis ......................................................................77 3.2.5 Basement Membranes ...........................................................77 3.3 Inflammation .........................................................................77

Clin Pharmacokinet 2005; 44 (1): 61-98<br />

REVIEW ARTICLE 0312-5963/05/0001-0061/$34.95/0<br />

© 2005 Adis Data Information BV. All rights reserved.<br />

Pharmacokinetics and<br />

Pharmacodynamics of Systemically<br />

Administered Glucocorticoids<br />

David Czock, Frieder Keller, Franz Maximilian Rasche and Ulla Häussler<br />

Division of Nephrology, University Hospital Ulm, Ulm, Germany<br />

Contents<br />

Abstract .....................................................................................62<br />

1. Pharmacokinetics .........................................................................63<br />

1.1 General ..............................................................................63<br />

1.1.1 Absorption and Distribution ......................................................63<br />

1.1.2 Metabolism and Excretion .......................................................66<br />

1.2 Selected Examples ....................................................................66<br />

1.2.1 Cortisol, Cortisone, and Hydrocortisone ...........................................66<br />

1.2.2 Prednisolone and Prednisone .....................................................66<br />

1.2.3 Methylprednisolone .............................................................67<br />

1.2.4 Dexamethasone ................................................................67<br />

1.3 Pharmacokinetic Drug-Drug Interactions ................................................68<br />

1.3.1 Influence of Other Drugs on Glucocorticoid Pharmacokinetics ......................68<br />

1.3.2 Influence of Glucocorticoids on the Pharmacokinetics of Other Drugs ................69<br />

1.3.3 Interactions Between Glucocorticoids and Ciclosporin, Tacrolimus or Sirolimus<br />

(Rapamycin) ...................................................................69<br />

1.4 Influence of Diseases on Glucocorticoid Pharmacokinetics ...............................70<br />

2. Genomic and Nongenomic Mechanisms of Glucocorticoid Effects ............................71<br />

2.1 Genomic Mechanisms ................................................................71<br />

2.1.1 Glucocorticoid Receptors ........................................................71<br />

2.1.2 Glucocorticoid Response Elements ...............................................71<br />

2.1.3 Transcription Factors (Nuclear Factor κB and Activator Protein 1) ....................73<br />

2.1.4 Post-Transcriptional and Translational Mechanisms ..................................73<br />

2.2 Nongenomic Mechanisms .............................................................73<br />

2.2.1 Specific Nongenomic Mechanisms ...............................................74<br />

2.2.2 Nonspecific Nongenomic Mechanisms ............................................75<br />

3. The Host Defence Response and the Effects of Glucocorticoids ...............................75<br />

3.1 Molecular Mechanisms ................................................................75<br />

3.1.1 Cytokines and Chemokines ......................................................76<br />

3.1.2 Inflammatory Enzymes ...........................................................76<br />

3.2 Cellular Mechanisms ..................................................................76<br />

3.2.1 Cell Trafficking and Adhesion Molecules ...........................................76<br />

3.2.2 T Cell Differentiation .............................................................76<br />

3.2.3 Cell Proliferation ................................................................77<br />

3.2.4 Apoptosis ......................................................................77<br />

3.2.5 Basement Membranes ...........................................................77<br />

3.3 Inflammation .........................................................................77


62 Czock et al.<br />

3.4 Immune Response ....................................................................78<br />

4. Adverse Effects of Glucocorticoids ..........................................................78<br />

5. Pharmacodynamics of Glucocorticoids .....................................................79<br />

5.1 Biomarker and Surrogate Endpoints .....................................................79<br />

5.2 Potency .............................................................................79<br />

5.3 Clinical Efficacy ......................................................................80<br />

5.4 Pharmacodynamic Drug-Drug Interactions ..............................................81<br />

6. Pharmacokinetic/Pharmacodynamic Models ................................................81<br />

6.1 Pharmacokinetic/Pharmacodynamic Analysis of Glucocorticoids .........................82<br />

6.2 In Vivo Potency ......................................................................82<br />

6.3 Selected Examples ....................................................................84<br />

6.3.1 Influence of Sex on Glucocorticoid Pharmacokinetic/Pharmacodynamic Properties . . . 84<br />

6.3.2 Influence of Age on Glucocorticoid Pharmacokinetic/Pharmacodynamic Properties 84<br />

6.3.3 Once- Versus Twice-Daily Glucocorticoid Administration ............................84<br />

6.3.4 Adverse Effects .................................................................85<br />

7. Clinical Aspects of Glucocorticoid Therapy ..................................................85<br />

7.1 Glucocorticoid Pulse Therapy ..........................................................85<br />

7.1.1 Emergency Treatment ...........................................................86<br />

7.1.2 Non-Emergency Treatment ......................................................86<br />

7.2 Diseases Not Treated with Pulse Therapy ................................................87<br />

7.3 Role of the Dosage Interval ............................................................87<br />

7.4 Variation in Clinical Response to Glucocorticoid Therapy .................................87<br />

8. Conclusions ..............................................................................87<br />

Abstract<br />

Glucocorticoids have pleiotropic effects that are used to treat diverse diseases<br />

such as asthma, rheumatoid arthritis, systemic lupus erythematosus and acute<br />

kidney transplant rejection. The most commonly used systemic glucocorticoids<br />

are hydrocortisone, prednisolone, methylprednisolone and dexamethasone. These<br />

glucocorticoids have good oral bioavailability and are eliminated mainly by<br />

hepatic metabolism and renal excretion of the metabolites. Plasma concentrations<br />

follow a biexponential pattern. Two-compartment models are used after intravenous<br />

administration, but one-compartment models are sufficient after oral administration.<br />

The effects of glucocorticoids are mediated by genomic and possibly nongenomic<br />

mechanisms. Genomic mechanisms include activation of the cytosolic<br />

glucocorticoid receptor that leads to activation or repression of protein synthesis,<br />

including cytokines, chemokines, inflammatory enzymes and adhesion molecules.<br />

Thus, inflammation and immune response mechanisms may be modified.<br />

Nongenomic mechanisms might play an additional role in glucocorticoid pulse<br />

therapy.<br />

Clinical efficacy depends on glucocorticoid pharmacokinetics and pharmacodynamics.<br />

Pharmacokinetic parameters such as the elimination half-life, and<br />

pharmacodynamic parameters such as the concentration producing the halfmaximal<br />

effect, determine the duration and intensity of glucocorticoid effects.<br />

The special contribution of either of these can be distinguished with pharmacokinetic/pharmacodynamic<br />

analysis. We performed simulations with a pharmacokinetic/pharmacodynamic<br />

model using T helper cell counts and endogenous cortisol<br />

© 2005 Adis Data Information BV. All rights reserved. Clin Pharmacokinet 2005; 44 (1)


Pharmacokinetics/Pharmacodynamics of Glucocorticoids 63<br />

as biomarkers for the effects of methylprednisolone. These simulations suggest<br />

that the clinical efficacy of low-dose glucocorticoid regimens might be increased<br />

with twice-daily glucocorticoid administration.<br />

mechanisms, pharmacodynamics, and clinical expe-<br />

rience with systemically administered glucocorti-<br />

coids.<br />

Glucocorticoids have pleiotropic effects and they<br />

are used frequently and intensively in clinical practice<br />

for many different indications. Their anti-inflammatory<br />

effect is used in inflammatory diseases<br />

(e.g. asthma and rheumatoid arthritis) and their immunosuppressive<br />

effect is used in autoimmune diseases<br />

(e.g. systemic lupus erythematosus [SLE]) as<br />

well as in organ transplantation (e.g. acute kidney<br />

transplant rejection). Clinically applied dosage regimens<br />

have been derived empirically and there is<br />

considerable variability between patients in clinical<br />

response to glucocorticoid treatment. The pharmacokinetics<br />

and pharmacodynamics of glucocorticoids<br />

have therefore been evaluated in many studies.<br />

1. Pharmacokinetics<br />

1.1 General<br />

The pharmacokinetic characteristics of the vari-<br />

ous glucocorticoids depend on their physicochemi-<br />

cal properties. [5] Glucocorticoids are lipophilic and<br />

are usually administered as prodrugs when given<br />

intravenously. Preparations include the hydrophilic<br />

phosphate and succinate esters of glucocorticoids,<br />

which are converted within 5–30 minutes to their<br />

active forms. [6-8] Small doses of glucocorticoids can<br />

also be administered as an alcoholic solution. [9-11]<br />

Generally, the indications for glucocorticoids can<br />

be divided into two categories. The first category<br />

includes emergency situations such as acute kidney<br />

transplant rejection or diffuse alveolar haemorrhage<br />

1.1.1 Absorption and Distribution<br />

due to autoimmune diseases. In these patients, high<br />

Glucocorticoids are well absorbed after oral addoses<br />

of glucocorticoids are usually administered<br />

ministration and have a bioavailability of 60–100%<br />

intravenously. The second category includes chronic<br />

(table I). [9,12-20] They have moderate protein binding<br />

diseases such as rheumatoid arthritis or the nephrotand<br />

a moderate apparent volume of distribution. The<br />

ic syndrome. In these patients, low-dose maintenpharmacokinetics<br />

of hydrocortisone (i.e. cortisol)<br />

ance therapy with glucocorticoids is given orally<br />

and prednisolone are nonlinear. Both bind to the<br />

using the lowest active dose in order to limit severe<br />

glycoprotein transcortin (i.e. corticosteroid binding<br />

long-term adverse events. What is the rationale beglobulin)<br />

and albumin. [21-23] Transcortin has a high<br />

hind these opposing practices, how might this be<br />

affinity and a low capacity for hydrocortisone and<br />

explained, and how might it be improved?<br />

prednisolone, whereas albumin has a low affinity<br />

Pharmacokinetic/pharmacodynamic modelling but high capacity. This leads to an increase in the<br />

allows simultaneous analysis of pharmacokinetics free glucocorticoid fraction once transcortin is satuand<br />

pharmacodynamics [1-4] and to distinguish their<br />

respective contribution to clinical efficacy. This<br />

might help to improve dosage regimens for glucocorticoids,<br />

such as intravenous pulse administration<br />

or low-dose administration of twice-daily dose fractions.<br />

This article aims to review and discuss the relationship<br />

between pharmacokinetics, molecular<br />

rated at concentrations of about 400 µg/L. Such<br />

concentrations are achieved after administration of<br />

hydrocortisone or prednisolone doses >20mg.<br />

Protein binding is biologically relevant, because<br />

only free drug can reach the biophase (i.e. the site of<br />

action) and interact with the receptor. Therefore,<br />

pharmacodynamic considerations have to include<br />

protein binding. Clinically, decreased protein bind-<br />

© 2005 Adis Data Information BV. All rights reserved. Clin Pharmacokinet 2005; 44 (1)


© 2005 Adis Data Information BV. All rights reserved. Clin Pharmacokinet 2005; 44 (1)<br />

Table I. Glucocorticoid pharmacokinetics after systemic administration. The pooled mean ± pooled standard deviation (see Appendix) and the range of the primary mean values<br />

(minimum–maximum) are given for cases where more than one study was found [3,6,7,9-20,22,24-71]<br />

Drug ROA Conc. F Cmax tmax t 1 /2 Vd Vd/F Vss CL CL/F PB fren ka c ke<br />

(%) (µg/L/1mg (h) (h) (L) b (L) b (L) b (L/h) b (L/h) b (%) (%) (h –1 ) (h –1 )<br />

dose) a<br />

Cortisol Total 94 0.35 ± 0.08<br />

(0.20–0.48)<br />

Hydrocortisone d IV Total 2.0 ± 0.3 27 ± 7 32 ± 4 18 ± 4 92 ± 2<br />

(1.7–2.1) (24–39) (92–93)<br />

PO Total 96 ± 20 15.3 ± 2.9 1.2 ± 0.4 1.8 ± 0.5 1.4 ± 0.9 0.43 ± 0.11<br />

Prednisolone IV Total<br />

phosphate<br />

Prednisolone after IV Total 88.3 ± 24.0 0.08 3.0 ± 0.4 45 ± 5 40 ± 9 9 ± 2 73 ± 4 18.0 ± 6.6 0.22 ± 0.03<br />

prednisolone (2.3–3.8) (39–50) (19–80) (5–16) (70–76) (10.3–24.4) (0.19–0.27)<br />

phosphate<br />

Free 2.2 ± 0.3 141 ± 44 48 ± 15 0.27 ± 0.06<br />

(1.7–2.7) (95–212) (20–63) (0.13–0.36)<br />

Prednisone after IV Total 3.7 ± 1.1 3.3 ± 1.4<br />

prednisolone (2.9–6.7) (2.1–5.3)<br />

phosphate<br />

Prednisolone IV Total 26 ± 13 50 ± 12 4.9 2.01 ± 0.48<br />

succinate (44–59)<br />

Prednisolone after IV Total 67 ± 23 12 ± 3 13.7 ± 4.2 0.19 ± 0.06<br />

prednisolone (57–81) (11–14) (12.1–14.9) (0.19–0.20)<br />

succinate<br />

Prednisolone PO Total 99 ± 8 18.1 ± 5.5 1.3 ± 0.7 3.2 ± 1.0 60 ± 28 67 ± 19 13 ± 4 86 ± 6 2.0 ± 2.5 0.26 ± 0.07<br />

(7.6–30.7) (0.9–1.6) (2.7–4.1) (30–132) (46–91) (8–23) (1.2–3.5) (0.19–0.32)<br />

Free 4.3 ± 2.8 1.5 ± 0.7 2.5 ± 1.0 302 ± 210 81 ± 42 1.1 0.31 ± 0.10<br />

(3.4–5.2) (1.4–1.5) (2.2–2.9) (254–350) (77–85) (0.27–0.36)<br />

Prednisone PO Total 84 ± 13 2.4 ± 1.1 2.6 ± 1.3 3.3 ± 1.3 2.4 ± 1.0<br />

(2.2–2.5) (2.5–2.8) (2.9–4.1) (1.6–2.0)<br />

Prednisolone after PO Total 79 ± 14 16.6 ± 4.8 1.9 ± 1.1 3.0 ± 0.8 43 ± 12 62 ± 9 12 ± 2 10.2 ± 3.5 3.9 ± 4.2 0.29 ± 0.11<br />

prednisone (62–99) (9.8–22.6) (1.3–3.0) (1.7–4.2) (31–48) (55–65) (7–14) (4.3–14.1) (1.3–9.7) (0.19–0.32)<br />

Free 53 ± 10 4.2 ± 1.1 2.0 ± 0.4 170 ± 54 72 ± 15 6.0 ± 6.2<br />

(1.6–2.6) (110–235) (38–97) (3.2–9.7)<br />

Methylprednisolone IV Total 0.25 ± 24 ± 6 27 ± 8.2 90 ± 27 9.2<br />

succinate 0.10 (23–25) (24.5–28.7) (82–97) (8.3–9.8)<br />

(0.07–0.37)<br />

Continued next page<br />

64 Czock et al.


© 2005 Adis Data Information BV. All rights reserved. Clin Pharmacokinet 2005; 44 (1)<br />

Table I. Contd<br />

Drug ROA Conc. F Cmax tmax t 1 /2 Vd Vd/F Vss CL CL/F PB fren ka c ke<br />

(%) (µg/L/1mg (h) (h) (L) b (L) b (L) b (L/h) b (L/h) b (%) (%) (h –1 ) (h –1 )<br />

dose) a<br />

Methylprednisolone IV Total 8.3 ± 3.3 0.8 2.4 ± 0.6 75 ± 16 90 ± 23 24 ± 7 78 ± 2 3.6 5.6 ± 3.5 0.28 ± 0.07<br />

after (6.3–10.7) (1.7–3.2) (55–95) (57–123) (13–33) (75–82) (3.3–3.9) (3.1–9.5) (0.19–0.41)<br />

methylprednisolone<br />

succinate<br />

Methylprednisolone IV Total 0.06 ± 6.1 ± 2.1 66 ± 21 1.2 ± 0.4 11.4 ± 2.5<br />

phosphate 0.01 (5.9–6.4) (53–78) (10.1–12.7)<br />

(0.06–0.07)<br />

Methylprednisolone IV Total 3.0 ± 1.7 76.3 ± 71.5 ± 24 ± 8 4.9 ± 1.5 0.28 ± 0.06<br />

after (3.0–3.1) 12.8 13.9 (23–24) (3.1–6.6) (0.23–0.33)<br />

methylprednisolone (60.8–91.8) (55.5–87.5)<br />

phosphate<br />

Methylprednisolone PO Total 88 ± 23 8.2 ± 2.4 2.1 ± 0.7 2.5 ± 1.2 100 ± 45 109 ± 32 26 ± 8 1.3 1.7 ± 0.5 0.27 ± 0.08<br />

(82–91) (4.6–10.4) (1.5–3.1) (1.6–3.4) (74–134) (96–125) (19–37) (0.23–0.31)<br />

Methylprednisolone 79 ± 3<br />

in vitro<br />

Methylprednisone 75 ± 3<br />

in vitro<br />

Dexamethasone IV Total 3.6 ± 28 ± 7 7.7 ± 1.6<br />

sodium phosphate 1.2<br />

Dexamethasone IV Total 90 10.5 ± 2.8 4.6 ± 1.2 65.7 ± 81.6 ± 12 ± 4 9.9 ± 18.6 0.21 ± 0.03<br />

after (10.2–10.8) (4.1–5.4) 17.3 16.6 (5–21) (4.0–12.4) (0.20–0.23)<br />

dexamethasone (27.0–98) (61.2–98)<br />

sodium phosphate<br />

Dexamethasone PO Total 76 ± 10 8.4 ± 3.6 1.5 4.0 ± 0.9 0.16<br />

(61–86) (1.0–2.0)<br />

Dexamethasone 75 ± 4<br />

in vitro<br />

a<br />

b<br />

c<br />

d<br />

Cmax was normalised to a glucocorticoid dose of 1mg. Multiply by 10 –6 /MW to convert from µg/L to mol/L. MW of hydrocortisone = 362.5Da, prednisolone = 360.5Da,<br />

prednisone = 358.4Da, methylprednisolone = 374.5Da and dexamethasone = 392.5Da.<br />

Volume and clearance parameters were normalised to 70kg bodyweight.<br />

Formation rate constant k f in the case of conversion after intravenous administration of a prodrug.<br />

Parameter only for low doses (= 20mg) of hydrocortisone available.<br />

CL = total clearance; CL/F = apparent clearance; C max = peak plasma concentration; Conc. = plasma concentration measured; F = fraction in % of the administered dose<br />

systemically available; Free = unbound to plasma components; f ren = renally excreted fraction in % of unchanged drug; IV = intravenous; ka =absorption rate constant; ke =<br />

elimination rate constant; MW = molecular weight; PB = plasma binding; PO = orally; ROA = route of administration; t max = time to reach C max; t 1 /2 = terminal half-life; Total =<br />

plasma bound and free; Vd = volume of distribution; Vd/F = apparent volume of distribution; V ss = volume of distribution at steady state.<br />

Pharmacokinetics/Pharmacodynamics of Glucocorticoids 65


66 Czock et al.<br />

HSD2 capacity and even leading to saturation of the<br />

mineralocorticoid receptor, would be expected to<br />

have enhanced mineralocorticoid effects when ad-<br />

ministered as two dose fractions, because the total<br />

time when mineralocorticoid receptors are occupied<br />

would be prolonged.<br />

ing due to low plasma albumin concentrations correlated<br />

with glucocorticoid adverse effects in prednisone<br />

therapy. [72] Generally, however, alterations in<br />

protein binding do not have much impact on drug<br />

action. [73]<br />

1.1.2 Metabolism and Excretion<br />

The renal excretion of unchanged glucocorticoids<br />

is only 1–20% (table I). [6,7,14,17,24-27] Glucocorticoid<br />

metabolism is a two-step process. Firstly,<br />

oxygen or hydrogen atoms are added then secondly,<br />

conjugation takes place (glucuronidation or sulpha-<br />

tion). Subsequently the kidney excretes the resulting<br />

hydrophilic inactive metabolites.<br />

Intracellular metabolism by 11β-hydroxysteroid<br />

dehydrogenase (11β-HSD) controls the availability<br />

of glucocorticoids for binding to the glucocorticoid<br />

and mineralocorticoid receptors. Type 1 dehydroge-<br />

nase (11β-HSD1) is widely distributed in glucocorticoid<br />

target tissues and has its highest activity in the<br />

liver. 11β-HSD1 acts mainly as a reductase, converting<br />

the inactive cortisone to the active cortisol.<br />

[74,75] Type 2 dehydrogenase (11β-HSD2) is<br />

found in mineralocorticoid target tissues (kidney,<br />

colon, salivary glands, placenta). 11β-HSD2 has a<br />

high affinity for endogenous cortisol and by oxida-<br />

tion, converting cortisol to cortisone, it protects the<br />

mineralocorticoid receptor from occupation by cortisol.<br />

[76] The activity of 11β-HSD2 varies depending<br />

on the type of glucocorticoid, [75,76] which explains to<br />

some extent the different mineralocorticoid activi-<br />

ties of different glucocorticoids.<br />

The undesired mineralocorticoid effects of glucocorticoid<br />

treatment should be pronounced when<br />

the capacity of 11β-HSD2 is exceeded. Therefore,<br />

we speculate that the mineralocorticoid effects of<br />

glucocorticoids might depend on the administration<br />

scheme. A low glucocorticoid dose leading to concentrations<br />

just above the protective capacity of<br />

11β-HSD2 would be expected to have reduced<br />

mineralocorticoid effects when administered as two<br />

dose fractions, because both concentration peaks<br />

would not exceed 11β-HSD2 capacity. In contrast,<br />

higher glucocorticoid doses, exceeding the 11β-<br />

1.2 Selected Examples<br />

1.2.1 Cortisol, Cortisone, and Hydrocortisone<br />

Cortisol is the active hormone produced by adrenal<br />

synthesis and secreted after stimulation by the<br />

pituitary hormone ACTH. Daily cortisol production<br />

is about 10mg in healthy volunteers [77] and can<br />

increase up to 400mg in conditions of severe<br />

stress. [78] Endogenous cortisol concentrations show<br />

a circadian pattern, with high concentrations in the<br />

morning between 6:00am and 9:00am (about 160<br />

µg/L at 8:00am in healthy volunteers) and low con-<br />

centrations in the evening between 8:00pm and<br />

2:00am. Cortisol is metabolised to the inactive cortisone<br />

and further to dihydrocortisone and tetrahy-<br />

drocortisone. Other metabolites include dihydrocor-<br />

tisol, 5α-dihydrocortisol, tetrahydrocortisol and 5α-<br />

tetrahydrocortisol. The biological activity of the latter<br />

metabolites is unclear. [79,80] After suppression of<br />

cortisol secretion by exogenous glucocorticoids,<br />

cortisol concentrations decline rapidly. Biexponen-<br />

tial functions are used to describe this cortisol de-<br />

cline after prednisolone [28,29] whereas monoexponential<br />

functions are used for the cortisol decline<br />

after methylprednisolone. [30,31]<br />

Hydrocortisone is chemically identical to cor-<br />

tisol, but this name is used in order to distinguish<br />

drug administration from endogenous production.<br />

Hydrocortisone is well absorbed after oral adminis-<br />

tration [9] and the disposition is biexponential. [81] The<br />

pharmacokinetic parameters of cortisol and hydro-<br />

cortisone are summarised in table I.<br />

1.2.2 Prednisolone and Prednisone<br />

Prednisolone (dehydrocortisol) is the active sub-<br />

stance, whereas the inactive prednisone (dehydro-<br />

cortisone) is activated by 11β-HSD1 to predniso-<br />

© 2005 Adis Data Information BV. All rights reserved. Clin Pharmacokinet 2005; 44 (1)


Pharmacokinetics/Pharmacodynamics of Glucocorticoids 67<br />

Table II. Dose-dependent pharmacokinetics of prednisolone, as shown for total concentration of prednisolone after oral administration. The<br />

pooled mean ± pooled standard deviation (see Appendix) and the range of the primary mean values (minimum–maximum) are given for<br />

cases where more than one study was found [19,32,35,36,53,60,67,70,71]<br />

Drug F (%) Cmax (µg/L/ tmax (h) t 1 /2 (h) Vd/F (L) b Vss (L) b CL/F (L/h) b ka (h –1 ) ke (h –1 )<br />

1mg dose) a<br />

Prednisolone 20.7 ± 6.5 1.3 ± 0.8 3.0 ± 1.0 42 ± 16 54 ± 16 10 ± 3 1.8 ± 2.1 0.28 ± 0.08<br />

(≤20mg) (16.2–30.7) (1.0–1.5) (2.7–3.7) (30–43) (46–63) (8–14) (1.4–3.5) (0.21–0.32)<br />

Prednisolone 16.3 ± 3.5 1.3 ± 0.6 3.0 ± 0.6 64 ± 13 91 ± 24 15 ± 4 2.2 ± 2.8 0.25 ± 0.06<br />

(25–60mg) (11.7–21.1) (0.9–1.6) (2.7–3.6) (51–74) (13–18) (1.2–3.2) (0.22–0.28)<br />

Prednisolone 99 ± 8 11.1 ± 6.0 1.5 ± 0.5 4.0 ± 1.4 c 98 ± 54 17 ± 7 0.19 ± 0.07<br />

(100mg) (7.6–12.3) (3.7–4.1) (87–132) (15–23)<br />

a C max was normalised to a prednisolone dose of 1mg.<br />

b Volume and clearance parameters were normalised to 70kg bodyweight.<br />

c There were no significant differences between the half-lives after high and low dose prednisolone in the primary studies.<br />

CL/F = apparent clearance; C max = peak plasma concentration; F = fraction in % of the administered dose systemically available; ka =<br />

absorption rate constant; k e = elimination rate constant; t 1 /2 = terminal half-life; tmax = time to reach Cmax; Vd/F = apparent volume of<br />

distribution; V ss = volume of distribution at steady state.<br />

lone. [76] Similar to cortisone/cortisol, there is inter- meters of prednisolone and prednisone are summarised<br />

conversion between both substances. The recycled<br />

in table I.<br />

proportion of prednisone has been estimated at<br />

1.2.3 Methylprednisolone<br />

76%. [14] Methylprednisolone (6α-methylprednisolone)<br />

The pharmacokinetics of prednisolone and pred- has no affinity for transcortin and binds only to<br />

nisone are complicated by dose-dependency due to albumin. [37,38] Accordingly, methylprednisolone<br />

nonlinear protein binding. [24,32] Protein binding of pharmacokinetics are linear, with no dose-dependprednisolone<br />

decreases nonlinearly from 95% to ency. [7,35] Methylprednisolone has many metabo-<br />

60–70%, while the concentration increases from 200 lites, including 20-carboxymethylprednisolone and<br />

µg/L to 800 µg/L when protein binding of prednisoterconversion<br />

6β-hydroxy-20α-hydroxymethylprednisolone. Inlone<br />

reaches a plateau. [18,24,33,34,82] In consequence, a<br />

between methylprednisolone and<br />

dose-dependent increase in the volume of distribuposition<br />

methylprednisone has been described. [37,39] The dis-<br />

tion (Vd) and drug clearance (CL) is observed at<br />

of methylprednisolone is biexponential. [7,17]<br />

doses over 20mg (table II). [19,32,35,83] However, the A two-compartment model is appropriate for intraelimination<br />

half-life remains constant [19,32] and the venous administration of very high doses. [17] A onedose<br />

dependencies of Vd and CL disappear when compartment model can be used with lower intrave-<br />

free prednisolone concentrations are measured. [18,35] nous doses [30,40] and oral administration. [36] The<br />

Prednisolone clearance decreases again only at very pharmacokinetic parameters of methylprednisolone<br />

high doses, which can be explained by saturation of are summarised in table I.<br />

elimination mechanisms. [27] The affinity of predni- 1.2.4 Dexamethasone<br />

sone for transcortin is 10-fold lower than that of Dexamethasone (9α-fluoro-16α-methylpredprednisolone.<br />

[82] Prednisolone metabolites include nisolone) has no affinity for transcortin and binds<br />

6β-hydroxyprednisolone and 20β-hydroxypred- only to albumin. [41] The pharmacokinetics of dexanisolone.<br />

[14,24] The disposition of prednisolone is methasone after intravenous administration are linebiexponential.<br />

A two-compartment model is appro- ar. [11,42] A second peak after intravenous administrapriate<br />

for intravenous administration. [27,29] One- and tion can be explained by enterohepatic recirculatwo-compartment<br />

models can be used with oral ad- tion. [20,43] The disposition of dexamethasone is<br />

ministration. [28,36,84] The pharmacokinetic para- biexponential [42,44] and a two-compartment model is<br />

© 2005 Adis Data Information BV. All rights reserved. Clin Pharmacokinet 2005; 44 (1)


68 Czock et al.<br />

Table III. Interactions of drugs and influence of diseases on glucocorticoid pharmacokinetics. The values are expressed as a percentage of<br />

the value without drug coadministration or disease (which was set at 100%) from the respective study [10,12,22,26,29,30,33,39,45-52,54,55,85,88-92]<br />

Drug Prednisolone Methylprednisolone Dexamethasone<br />

CL/F (%) t 1 /2 (%) CL/F (%) t 1 /2 (%) CL/F (%) t 1 /2 (%)<br />

Phenobarbital 140 80 300 45<br />

(phenobarbitone)<br />

Carbamazepine 180 65 440 45<br />

Phenytoin 180 70 580 30<br />

Rifampicin (rifampin) 140 60<br />

Ketoconazole Unchanged Unchanged 40–50<br />

Itraconazole Unchanged Unchanged 170 30 300<br />

Diltiazem 83 115 67 145<br />

Troleandomycin 190<br />

Erythromycin<br />

Prolonged<br />

Clarithromycin Unchanged Unchanged 35 230<br />

Ciclosporin Decreased Unchanged Unchanged<br />

Sirolimus (rapamycin) 130<br />

Grapefruit juice Unchanged 130<br />

Disease<br />

Renal failure 60 150 Unchanged Unchanged 165 70<br />

Chronic liver disease Reduced Unchanged Unchanged 65 170<br />

CL/F = apparent clearance; t 1 /2 = terminal half-life.<br />

appropriate for intravenous dexamethasone. [43,44] prednisolone elimination. In another study, the loss<br />

The pharmacokinetic parameters of dexamethasone of kidney transplant function was associated with<br />

are summarised in table I. rifampicin treatment. [87]<br />

1.3 Pharmacokinetic Drug-Drug Interactions<br />

Coadministration of cytochrome P450 (CYP)<br />

3A4 inhibitors (e.g. ketoconazole, clarithromycin)<br />

decreases the clearance and increases the half-life of<br />

1.3.1 Influence of Other Drugs on<br />

Glucocorticoid Pharmacokinetics<br />

methylprednisolone and dexamethasone, whereas<br />

Coadministration of enzyme inducers (e.g. barbi- prednisolone is usually not affected (table<br />

turates, carbamazepine, phenytoin, rifampicin [riticoids<br />

III). [12,26,29,30,46-49] Decreased clearance of glucocor-<br />

fampin]) increases the clearance and decreases the<br />

can lead to increased effects (biomarker:<br />

half-life of prednisolone and methylprednisolone lymphocytes, cortisol). [30,39,47,50,51,93-95] Clinically,<br />

(table III). [85] A study of the time course of induction macrolides have been used as glucocorticoid-sparwith<br />

rifampicin showed that the changes in pharmaasthma.<br />

ing agents in patients with glucocorticoid-dependent<br />

cokinetics of prednisolone were maximal 2 weeks<br />

[96,97] However, short-term administration of<br />

after the start, and were normal again 2 weeks after macrolides does not need a dose reduction of gluco-<br />

the end of rifampicin therapy. [45] The clinical relepharmacokinetics<br />

corticoids. Cimetidine had no influence on the<br />

vance of such interactions was demonstrated in paprednisolone<br />

of prednisolone [83,98] or methyl-<br />

tients with kidney transplants treated with prednicreased<br />

in one patient. [99] Grapefruit juice intients<br />

sone. A lower kidney transplant survival rate was<br />

the half-life of oral methylprednisolone [52]<br />

found in those patients receiving concomitant anticonvulsants<br />

but did not affect prednisolone. [88]<br />

(phenobarbital [phenobarbitone]/ Users of oral contraceptives (OCs) have higher<br />

diphenylhydantoin). [86] This can be explained by a concentrations [53] and a lower clearance [34] of predreduced<br />

immunosuppressive effect due to increased nisolone, which has been explained by increased<br />

© 2005 Adis Data Information BV. All rights reserved. Clin Pharmacokinet 2005; 44 (1)


Pharmacokinetics/Pharmacodynamics of Glucocorticoids 69<br />

transcortin levels. [34] Methylprednisolone has a low- nosyltransferase induction after methylprednisolone<br />

er clearance in OC users, [40] which has been ex- withdrawal could explain increased mycophenolate<br />

plained by inhibition of oxidative processes by OCs. concentrations. Induction of glucuronosyltransfer-<br />

This lower clearance of methylprednisolone, how- ase expression by glucocorticoids was demonstrated<br />

ever, was associated with mixed changes in in vitro. [113]<br />

glucocorticoid effects, and it was concluded that a<br />

change in methylprednisolone dose is not necessary 1.3.3 Interactions Between Glucocorticoids and<br />

in OC users. [40]<br />

Ciclosporin, Tacrolimus or Sirolimus (Rapamycin)<br />

In vitro and animal studies suggest that cortisol, The interactions between ciclosporin and glucoprednisolone,<br />

methylprednisolone, and dexamethconflicting<br />

corticoids are complicated and there are studies with<br />

asone are substrates of P-glycoprotein. [100-102] Therecoids<br />

results. Both ciclosporin and glucocorti-<br />

fore, coadministration of P-glycoprotein inhibitors<br />

are substrates of CYP [12,30] and possibly P-<br />

could increase glucocorticoid absorption and might glycoprotein. [100] Ciclosporin inhibits P-glycoproteaffect<br />

glucocorticoid distribution. Oral coadminprotein<br />

in, whereas glucocorticoids might induce P-glyco-<br />

istration of valspodar increased the area under the<br />

and CYP. [106,110]<br />

curve (AUC) of dexamethasone 1.24 times in Ciclosporin affects the pharmacokinetics of predhealthy<br />

volunteers, although it is most likely that nisolone. The AUC of prednisolone after oral predthis<br />

is not clinically relevant. [103]<br />

nisone was higher, [89] and the oral clearance of pred-<br />

Smoking did not affect the pharmacokinetics of nisolone was lower [90] in patients receiving ciprednisolone<br />

or dexamethasone. [104]<br />

closporin. In addition, the oral clearance of<br />

prednisolone increased in patients who were<br />

1.3.2 Influence of Glucocorticoids on the<br />

switched to a ciclosporin-free regimen and de-<br />

Pharmacokinetics of Other Drugs<br />

creased when ciclosporin was added. [90] This might<br />

Glucocorticoids affect the pharmacokinetics of be explained at least partially by increased absorpother<br />

drugs by enzyme induction (e.g. of cyto- tion of prednisolone due to P-glycoprotein inhibichromes,<br />

[105] P-glycoprotein or glucuronosyltrans- tion by ciclosporin. In contrast, another study did<br />

ferase). Dexamethasone induced CYP3A4 at high not find changes in the pharmacokinetics of prednisdoses<br />

(16–24 mg/day) [106] but not at low doses (1.5 olone with ciclosporin coadministration. [114] Howmg/day).<br />

[107] Methylprednisolone at a daily dose of ever, the latter study was performed in patients only<br />

8mg did not induce CYP3A4 in healthy volun- 1 month after transplantation and it is possible that<br />

teers. [108] Autoinduction has been used to explain a high doses of glucocorticoids in the early posttime-dependent<br />

increase in the prednisolone [35] and transplantation period induced P-glycoprotein and<br />

methylprednisolone [109] clearance. P-glycoprotein CYP which counterbalanced the inhibitory effects<br />

might be induced by lower glucocorticoid doses. of ciclosporin. Another observation is that oral pred-<br />

Dexamethasone induced P-glycoprotein at low nisolone clearance was lower at 3–6 months than at<br />

doses and CYP at high doses, as shown in rats. [110]


70 Czock et al.<br />

Kidney transplant patients on ciclosporin had an mained unchanged. [10,33] The protein binding of<br />

intravenous clearance of methylprednisolone that prednisolone decreased and thus the free fraction<br />

was similar to the clearance in healthy volunteers increased in renal failure. [33] Methylprednisolone<br />

from the literature. [54] Sirolimus (rapamycin) re- pharmacokinetics were unchanged in patients with<br />

duced the oral clearance of prednisolone after pred- renal failure. Only the conversion of the prodrug<br />

nisone and a negative correlation between sirolimus methylprednisolone succinate to methylpred-<br />

AUC and prednisolone clearance was observed. [91] nisolone was slower. [55] In contrast, dexamethasone<br />

Methylprednisolone affects ciclosporin pharma- clearance was increased and its half-life decreased<br />

cokinetics. The clearance of intravenous ciclosporin in renal failure (table III). [10] This can be explained<br />

was increased after methylprednisolone pulse ther- by decreased protein binding of dexamethasone to<br />

apy (250 mg/day for 3 days), [116] which can be albumin in uraemia. [41] Similarly, the binding of<br />

explained by CYP induction. In contrast, another cortisol to albumin is reduced in uraemia, [23] wherestudy<br />

found higher ciclosporin plasma levels after as binding of cortisol to transcortin remains unmethylprednisolone<br />

(250–500 mg/day). [117] How- changed. [22] Cortisol metabolites accumulate in reever,<br />

in the latter study a nonspecific RIA was used nal failure and are removed insufficiently during<br />

and thus the contribution of ciclosporin metabolites haemodialysis. [79] The fraction of prednisolone reto<br />

the measured plasma levels is unclear. Ci- moved during a 5-hour haemodialysis was<br />

closporin pharmacokinetics were unaffected by 7–17.5%. [121] The fraction of methylprednisolone<br />

withdrawal of low-dose prednisolone for 24 hours in removed during haemodialysis can be estimated to<br />

liver transplant recipients. [118] However, enzyme in- be


Pharmacokinetics/Pharmacodynamics of Glucocorticoids 71<br />

There are a number of reports for other diseases<br />

that may affect glucocorticoid pharmacokinetics.<br />

Patients with the nephrotic syndrome had decreased<br />

total but normal unbound prednisolone concentra-<br />

tions. [123] Patients with inflammatory bowel disease<br />

have variable absorption, [83,124] and fractional bind-<br />

ing of prednisolone to plasma proteins was decreased<br />

in active disease. [125] Hyperthyroid patients<br />

had decreased prednisolone concentrations. [83] Patients<br />

with cystic fibrosis had a higher volume of<br />

distribution and normal half-life of oral predniso-<br />

lone, [126] whereas prednisolone pharmacokinetics in<br />

steroid-dependent asthmatics were unchanged compared<br />

with healthy individuals. [127] Children with<br />

acute lymphoblastic leukaemia had a lower clear-<br />

ance and longer half-life. [128] Patients with acute<br />

spinal cord injury had a lower systemic clearance of<br />

methylprednisolone. [129]<br />

2. Genomic and Nongenomic<br />

Mechanisms of Glucocorticoid Effects<br />

Glucocorticoids inhibit many events of the inflammatory<br />

and immune response by various mechanisms<br />

(table IV). [130,131] Most effects of glucocorticoids<br />

are mediated by glucocorticoid receptors, but<br />

possibly not all. Basically, glucocorticoid mechanisms<br />

can be divided into genomic and nongenomic<br />

mechanisms. Whereas the role of genomic mechanisms<br />

is well established, the importance of nongenomic<br />

mechanisms is still unclear.<br />

2.1 Genomic Mechanisms<br />

The genomic effects of glucocorticoids are characterised<br />

by a slow onset and slow dissipation of<br />

the response, due to the time-consuming process of<br />

mRNA transcription and translation (figure 1). The<br />

start of mRNA induction for some proteins was seen<br />

after 15 minutes in rat thymic lymphocytes. [132] Protein<br />

levels can be affected after 30 minutes and<br />

effects on tissue or organ levels need hours to days.<br />

It can be estimated that 100 to 1000 genes are up- or<br />

down-regulated in a cell-type specific way. [133-135]<br />

2.1.1 Glucocorticoid Receptors<br />

The glucocorticoid receptor (GRα), a member of<br />

a superfamily of ligand-regulated nuclear receptors,<br />

consists of a ligand-binding domain, a DNA-binding<br />

domain, and two activation-functional domains. The<br />

inactive glucocorticoid receptor is retained in the<br />

cytoplasm within a multiprotein complex containing<br />

two heat shock protein molecules (HSP90). Addi-<br />

tional molecules of this complex include the immunophyllins,<br />

HSP70, p23, and src. [136-138] The gluco-<br />

corticoid receptor isoform GRβ, a splice variant,<br />

does not bind glucocorticoids and its relevance in<br />

glucocorticoid resistance is controversial.<br />

Glucocorticoids bind to the ligand-binding do-<br />

main of the glucocorticoid receptor GRα after entering<br />

the cell by passive diffusion through the cell<br />

membrane. This binding leads to a conformational<br />

change in the glucocorticoid receptor, which in turn<br />

leads to dissociation of the multiprotein complex<br />

(figure 1). The glucocorticoid receptor-glucocorticoid<br />

(GR/GC) complex is then localised to the nucleus<br />

via the nuclear pore complex. Dimerisation<br />

with formation of glucocorticoid receptor homod-<br />

imers is necessary for DNA interaction with gluco-<br />

corticoid response elements (GREs) and subsequent<br />

activation or repression of transcription (figure<br />

1). [139,140] The transactivation potency of the GR/GC<br />

complex is regulated intracellularly by coactivators<br />

and corepressors, [141,142] and possibly by glucocorti-<br />

coid receptor phosphorylation. [143,144] In addition,<br />

monomeric glucocorticoid receptors inhibit the<br />

proinflammatory transcription factors nuclear factor<br />

κB (NF-κB) [figure 1] and activator protein 1 (AP-<br />

1), independent of DNA binding. The importance of<br />

the latter mechanism is underlined by the fact that<br />

many inflammatory genes that are repressed by glu-<br />

cocorticoids do not contain a negative GRE.<br />

2.1.2 Glucocorticoid Response Elements<br />

The GREs in gene promoters are characterised by<br />

the nonpalindromic consensus sequence 5′-XXTA-<br />

CAXXXTGTTCT-3′ containing two binding sites<br />

for the glucocorticoid receptor homodimer. After<br />

DNA association, the GR/GC homodimer (together<br />

© 2005 Adis Data Information BV. All rights reserved. Clin Pharmacokinet 2005; 44 (1)


72 Czock et al.<br />

Table IV. Molecular mechanisms of glucocorticoid effects<br />

Mechanisms Molecular effects Cellular effects<br />

Genomic mechanisms<br />

Transcriptional<br />

transactivation Interactions of GR with GREs Anti-inflammatory/immunosuppressive effects<br />

Interactions of GR with transcription factors induction of anti-inflammatory cytokines<br />

(→ CBP → acetylation of core histones (e.g. IL-10, TGFβ)<br />

→ increased gene transcription)<br />

induction of cytokine receptors<br />

(e.g. IL-1RII, IL-10R, TGFβR)<br />

induction of proapoptotic factors<br />

Metabolic effects<br />

induction of PEPCK, TAT (gluconeogenesis)<br />

mobilisation of amino and fatty acids<br />

Antiproliferative effects on non-immune cells<br />

induction of p21CIP1 (e.g. renal mesangial cells)<br />

induction of MKP-1 (e.g. osteoblasts)<br />

Other effects<br />

antiapoptotic effect (e.g. induction of c-IAP2)<br />

up-regulation of β 2-receptors<br />

transrepression Interactions of GR with nGREs Anti-inflammatory/immunosuppressive effects<br />

Interactions of GR with transcription factors suppression of:<br />

inhibition of AP-1, NF-κB<br />

cytokines (e.g. IL-1, IL-2, IL-6, IL-12, IFNγ)<br />

→ CBP associated HAT activity ↓<br />

chemokines (e.g. MCP-1, IL-8, eotaxin)<br />

→ inhibition of histone acetylation<br />

receptor expression (e.g. IL-2R)<br />

→ decreased gene transcription<br />

adhesion molecules<br />

direct (e.g. ICAM-1, E-selectin)<br />

indirect via cytokine/chemokine suppression<br />

(e.g. of IL-1β, TNFα)<br />

inflammatory enzymes (e.g. COX-2, cPLA 2, iNOS)<br />

T cell proliferation (e.g. via IL-2↓)<br />

Other effects<br />

suppression of the hypothalamic-pituitary-adrenal axis<br />

suppression of osteocalcin<br />

suppression of matrix metalloproteinase<br />

Post-transcriptional Modification of mRNA stability Suppression of COX-2, MCP-1, iNOS<br />

(shortening of the poly(A) tail)<br />

Translational<br />

Post-translational<br />

Suppression of ribosomal proteins and<br />

translation initiation factors<br />

Protein processing, secretion<br />

Nongenomic mechanisms<br />

Specific<br />

classical GR Cytosolic interactions (possibly via cPLA 2 inhibition (via src/annexin-1)<br />

components of the GR-multiprotein complex) Tertiary CAM structure (possibly via annexin-1)<br />

nonclassical GR Interaction with membrane GR May induce apoptosis<br />

Interaction with other receptors<br />

May induce IP 3, Ca2+, protein kinase C, cAMP, MAPK<br />

Continued next page<br />

© 2005 Adis Data Information BV. All rights reserved. Clin Pharmacokinet 2005; 44 (1)


Pharmacokinetics/Pharmacodynamics of Glucocorticoids 73<br />

Table IV. Contd<br />

Mechanisms Molecular effects Cellular effects<br />

Nonspecific<br />

Glucocorticoid dissolves in membranes and may<br />

alter<br />

→ physicochemical membrane properties<br />

(fluidity, ‘membrane stabilisation’)<br />

→ activity of membrane associated proteins<br />

AP-1 = activator protein 1; CAM = cellular adhesion molecule; cAMP = cyclic adenosine monophosphate; CBP = CREB binding protein; c-<br />

IAP2 = cellular inhibitor of apoptosis; COX-2 = cyclo-oxygenase 2; cPLA 2 = cytosolic phospholipase A 2; GR = glucocorticoid receptor; GRE<br />

= glucocorticoid response element; HAT = histone acetylase; ICAM = intercellular adhesion molecule; IFN = interferon; IL = interleukin;<br />

iNOS = inducible nitric oxide synthase; IP 3 = inositol-1,4,5-trisphosphate; MAPK = mitogen activated protein kinase; MCP-1 = monocyte<br />

chemoattractant protein 1; MKP-1 = MAP kinase phosphatase 1; mRNA = messenger RNA; NF-κB = nuclear factor κB; nGRE = negative<br />

GRE; PEPCK = phosphoenolpyruvate carboxykinase; TAT = tyrosine aminotransferase; TGFβ = transforming growth factor-β; TNF =<br />

tumour necrosis factor.<br />

with other cofactors) binds the transcriptional coac- proinflammatory p65-CBP HAT complex leads to<br />

tivator CREB binding protein (CBP)/p300 which in deacetylation of specific histone residues, DNA<br />

turn binds the basal transcription factor apparatus folding, and thus to reduced transcription or silencand<br />

starts gene transcription. DNA transcription, ing of proinflammatory genes. [150,151] Induction of<br />

once started, is independent of the further presence IκB or HDAC might play an additional role as a<br />

of the GR/GC complex (‘hit and run’ princi- long-term anti-inflammatory effect of glucocortiple).<br />

[145,146] coids. [130,152]<br />

Acetylation and deacetylation of specific histone<br />

residues regulates histone-chromatin interaction and 2.1.4 Post-Transcriptional and<br />

Translational Mechanisms<br />

thus the accessibility of genes for transcription factors.<br />

CBP has intrinsic histone acetylase (HAT) Protein synthesis depends on the stability and<br />

activity and thus can promote anti-inflammatory half-life of mRNA, which is regulated by the length<br />

gene transcription via histone acetylation and concorticoids<br />

can act by destabilising the mRNA of<br />

of its poly(A) tail and other mechanisms. [130] Gluco-<br />

secutive unfolding of the DNA.<br />

proinflammatory proteins. [153,154] Additionally, glucocorticoids<br />

might act at the translational level by<br />

2.1.3 Transcription Factors (Nuclear Factor κB and<br />

Activator Protein 1)<br />

suppression of ribosomal proteins and translation<br />

The inflammatory response of tissue cells, stimulated,<br />

initiation factors.<br />

for example, by tumour necrosis factor<br />

(TNFα) or interleukin (IL)-1β, is mediated intracel- 2.2 Nongenomic Mechanisms<br />

lularly by the proinflammatory transcription factors<br />

NF-κB (e.g. p65-p50 heterodimer) [figure 1] and Nongenomic mechanisms are characterised by a<br />

AP-1 (Fos-Jun heterodimer). [147,148] In later phases rapid onset of effect (


74 Czock et al.<br />

Glucocorticoid<br />

Cytokine receptor<br />

Glucocorticoid receptor –<br />

HSP90 complex<br />

NF<br />

IκB<br />

−κB<br />

Specific nongenomic<br />

effects<br />

NF<br />

−κB<br />

Anti-inflammatory genes<br />

NF −κB GRE<br />

Proinflammatory<br />

genes<br />

Fig. 1. Genomic mechanisms of glucocorticoids in human cells. Mechanisms include transactivation via binding to a glucocorticoid response<br />

element (GRE) in the promoter region of a gene (e.g. interleukin-10 gene) and transrepression via inhibition of the transcription factor NFκB.<br />

HSP90 = heat shock protein 90; IκB = inhibitor of NF-κB; NF = nuclear factor.<br />

Nongenomic effects have been used to explain<br />

the increased clinical effect of pulse therapy with<br />

glucocorticoid doses >250mg. Genomic mechanisms<br />

are not sufficient to explain such effects, as it<br />

has been estimated that all glucocorticoid receptors<br />

are occupied after prednisolone 100–200mg. [157] Ef-<br />

fects associated with nongenomic mechanisms can<br />

lead to different in vitro drug potencies compared<br />

with genomically mediated effects. [160,161]<br />

2.2.1 Specific Nongenomic Mechanisms<br />

Specific nongenomic mechanisms are mediated<br />

via the classical glucocorticoid receptor (figure 1) or<br />

nonclassical glucocorticoid receptors. [159] For exam-<br />

ple, a fast (


Pharmacokinetics/Pharmacodynamics of Glucocorticoids 75<br />

liberation of annexin 1 (formerly called lipocortin lysosomal membranes might be another mechanism<br />

1), which interferes with intracellular signal trans- (figure 2). [165]<br />

duction and eventually inhibits cPLA activation. [136]<br />

Experiments with cells from an annexin 1 knockout 3. The Host Defence Response and the<br />

mouse support these findings. [162]<br />

Effects of Glucocorticoids<br />

2.2.2 Nonspecific Nongenomic Mechanisms<br />

In order to understand the multiple effects of<br />

Nonspecific nongenomic mechanisms are due to<br />

glucocorticoids, a review of the host defence resdirect<br />

interaction of glucocorticoids with cell meminflammation,<br />

the immune response, coagulation,<br />

ponse is needed. The host defence response includes<br />

branes. It has been suggested that the lipophilic<br />

steroids physically dissolve into lipid membranes<br />

tissue repair and activation of the hypothalamic-<br />

and modify physicochemical membrane properties,<br />

pituitary-adrenal axis.<br />

which in turn affect the activity of membrane-associated<br />

proteins. [163,164] Suppression of cellular energy<br />

3.1 Molecular Mechanisms<br />

metabolism leading to reduced cell function On the molecular level, inflammation and the<br />

could be a consequence. [157] Stabilisation of immune response are mediated by cytokines (e.g.<br />

Glucocorticoid<br />

?<br />

Basement<br />

membrane<br />

?<br />

Cell membrane<br />

Lysosome<br />

Fig. 2. Nonspecific nongenomic mechanisms of glucocorticoids. These might affect membranes by physicochemical mechanisms. Involved<br />

membranes include cell membranes, lysosomal membranes and possibly basement membranes.<br />

© 2005 Adis Data Information BV. All rights reserved. Clin Pharmacokinet 2005; 44 (1)


76 Czock et al.<br />

IL-1 to IL-6, IL-11 to IL-13, TNFα, interferon iNOS, which generates nitric oxide, is involved<br />

[IFN]γ, macrophage migration inhibitory factor in vasodilatation at the site of inflammation. Gluco-<br />

[MIF]), chemokines (e.g. IL-8, monocyte chemoat- corticoids can suppress iNOS expression by trantractant<br />

protein [MCP]-1, macrophage inflamma- scriptional and post-transcriptional mechanisms as<br />

tory protein [MIP]1α, regulated upon activation nor- shown in vitro. [169,170] Glucocorticoid treatment remal<br />

T cell expressed and secreted [RANTES], duced nitric oxide levels in exhaled air in patients<br />

eotaxin), kinins and kinin receptors, adhesion mole- with pulmonary sarcoidosis or cystic fibrosis. [171,172]<br />

cules, and inflammatory enzymes (e.g. inducible<br />

nitric oxide synthase [iNOS], cyclo-oxygenase<br />

[COX]-2).<br />

3.2 Cellular Mechanisms<br />

3.1.1 Cytokines and Chemokines<br />

3.2.1 Cell Trafficking and Adhesion Molecules<br />

Cytokines, extracellular signalling proteins that The inflammatory process depends on migration<br />

induce cellular responses, affect protein production, of inflammatory and anti-inflammatory immune<br />

antigen expression and proliferation. Glucocorti- cells to the site of inflammation (cell trafficking).<br />

coids act by suppression of proinflammatory cytochemokines<br />

Activation of immune cells is mediated by<br />

kines (e.g. IL-1β, IFNα), induction of decoy recepcells<br />

(e.g. MCP-1, IL-8). Exit of immune<br />

tors that trap proinflammatory cytokines (e.g.<br />

from the blood vessels is mediated by cellular<br />

IL-1RII), induction of anti-inflammatory cytokines adhesion molecules (CAMs). [173]<br />

(e.g. transforming growth factor [TGF]-β, IL-10), Glucocorticoids can reduce the recruitment of<br />

and induction of anti-inflammatory cytokine recepsion<br />

immune cells to the site of inflammation by repres-<br />

tors (e.g. TGFβR, IL-10R). [135,166]<br />

of adhesion molecules, either directly as shown<br />

Chemokines, extracellular signalling proteins in vitro [174,175] or indirectly via suppression of prointhat<br />

affect cellular migration, can be suppressed flammatory cytokines and transcription factors. In<br />

(e.g. MCP-1 = CCL2, IL-8 = CXCL8, MIP-1β = addition, glucocorticoids can induce rapid changes<br />

CCL4, eotaxin = CCL11) or induced (e.g. internongenomic<br />

in the surface distribution of CAMs, possibly by a<br />

feron-inducible protein IP-10 = CXCL10,<br />

mechanism. [173,176] Glucocorticoid-in-<br />

fractalkine = CX3CL1) by glucocorticoids. duced granulocytosis can be explained by limited<br />

neutrophil emigration from the blood and neutrophil<br />

3.1.2 Inflammatory Enzymes mobilisation from the bone marrow.<br />

Arachidonic acid, an important mediator of the<br />

inflammatory response, is produced by phospho- 3.2.2 T Cell Differentiation<br />

lipase A2 (PLA2). Arachidonic acid is metabolised T Cell subsets include CD4+ helper (Th) cells,<br />

by COX, by thromboxane synthase, and by lipox- CD8+ cytotoxic T (Tc) cells, CD4+CD25+ natural<br />

ygenase, but arachidonic acid may also have direct regulatory T (Tr) cells and adaptive/inducible regueffects<br />

itself as a second messenger. [167] COX-2, the latory T (Treg1) cells [177,178] and CD8+CD28- supinducible<br />

COX, is induced by inflammatory and pressor T cells. [179] Naive T helper (Th0) cells differmitogenic<br />

stimuli.<br />

entiate to Th1 or Th2 cells depending on the stimu-<br />

Glucocorticoids suppress cytosolic cPLA2 and lus. [180] Regulatory T cells suppress immune<br />

COX-2 expression via genomic mechanisms. [168] responsiveness and induce tolerance. [181] Natural Tr<br />

Additionally, a specific nongenomic mechanism has cells express membrane-bound TGFβ that can delivbeen<br />

suggested for suppression of cPLA 2 ac- er signals to target cells via a contact-dependent<br />

tivity. [136] As a consequence, the production of process. [182] Adaptive Treg1 cells produce IL-10 and<br />

arachidonic acid and its metabolites is decreased. TGFβ. [178]<br />

© 2005 Adis Data Information BV. All rights reserved. Clin Pharmacokinet 2005; 44 (1)


Pharmacokinetics/Pharmacodynamics of Glucocorticoids 77<br />

Glucocorticoids inhibit the costimulatory CD40- philic inflammation prevails. In contrast, glucocortiligand<br />

on lymphocytes [135] and costimulatory mole- coids decrease the rate of apoptosis in neutrophilic<br />

cules on dendritic cells (e.g. CD40, CD86). [183] granulocytes, which could explain the limited clin-<br />

Eventually, dendritic cells are switched to produce ical efficacy of long-term glucocorticoid treatment<br />

IL-10 instead of IL-12, [184] which in turn limits in chronic obstructive pulmonary disease, where<br />

differentiation of Th0 to Th1 cells. In addition, neutrophilic inflammation prevails. [193,194]<br />

glucocorticoids might induce differentiation of Th0 Glucocorticoids can induce apoptosis by several<br />

to Treg1 cells. [185,186] Up-regulation of TGFβ recep- mechanisms. Firstly, cytokines that represent survitors<br />

on lymphocytes might enhance the action of val factors for immune cells are down-regulated by<br />

regulatory T cells.<br />

glucocorticoids, which can lead indirectly to<br />

apoptosis. Secondly, glucocorticoids might induce<br />

3.2.3 Cell Proliferation proapoptotic factors. Thirdly, glucocorticoids might<br />

Glucocorticoids affect the proliferation of inhibit NF-κB mediated antiapoptotic mechanimmune<br />

and nonimmune cells. Indirect glucocorti- isms. [195]<br />

coid effects include inhibition of T cell growth fac- Antiapoptotic mechanisms of glucocorticoids intor<br />

production (e.g. IL-2), which leads to reduced T clude induction of receptors for antiapoptotic sigcell<br />

proliferation. Direct glucocorticoid effects in- nals [196] and induction of intracellular antiapoptotic<br />

clude transcription of the protein p21CIP1, an inhibi- factors (e.g. the cellular inhibitor of apoptosis c-<br />

tor of cyclin-dependent kinase, which leads to cell IAP2). [197]<br />

cycle arrest as shown in renal mesangial cells. [187]<br />

Importantly, glucocorticoids have to be studied over 3.2.5 Basement Membranes<br />

a wide concentration range. Low to medium concentein-free<br />

The renal glomerulus normally produces a pro-<br />

trations of prednisolone (10 –9 to 10 –6 mol/L) stimupermeability<br />

filtrate, but in the nephrotic syndrome the<br />

lated intestinal epithelial cell proliferation, whereas<br />

of the glomerular capillary wall for<br />

very high concentrations (10 –4 mol/L) inhibited increase<br />

macromolecules is increased. Glucocorticoids detestinal<br />

epithelial cell proliferation. [188]<br />

proteoglycan synthesis in glomerular epi-<br />

thelial and mesangial cells by genomic mechan-<br />

3.2.4 Apoptosis isms. [198,199] Glucocorticoids inhibit the release of<br />

Glucocorticoids have apoptotic or antiapoptotic matrix metalloproteinases, as shown in alveolar<br />

effects, depending on the cell type. Apoptosis of macrophages. [200] In addition, nonspecific glucocorimmune<br />

cells (e.g. lymphocytes) leads to attenua- ticoid effects on basement membranes might be<br />

tion of the inflammatory and immune response, assumed.<br />

whereas antiapoptotic effects could protect resident<br />

cells (e.g. epithelial cells) of the inflamed tissue. [189]<br />

3.3 Inflammation<br />

Apoptosis of T cells could be an important mechanism<br />

of intravenous glucocorticoid pulse therapy. Acute inflammation is characterised by four con-<br />

Administration of methylprednisolone 500–1000mg secutive phases. These are exudation, local infiltrainduced<br />

apoptosis of T helper cells in patients with tion by neutrophils, apoptosis of neutrophils, and<br />

autoimmune diseases. [190,191] The degree of T cell local infiltration of mononuclear cells. Exudation is<br />

apoptosis was dependent on the glucocorticoid dose due to increased vascular permeability of capillaries<br />

in an animal model. [192]<br />

and venules. Vasoactive factors (e.g. nitric oxide,<br />

Glucocorticoids increase the rate of apoptosis in prostacyclin) and reduced adrenergic receptor aceosinophilic<br />

granulocytes in vitro, which could be tivity lead to dilatation of arterioles producing local<br />

important in the treatment of asthma, where eosino- erythema and heat. Infiltration by neutrophils is<br />

© 2005 Adis Data Information BV. All rights reserved. Clin Pharmacokinet 2005; 44 (1)


78 Czock et al.<br />

mediated by adhesion molecules and chemokines. regulated in acute rejection [205,209] could play a role.<br />

Apoptosis and phagocytosis of neutrophils by Finally, inhibition of the effector mechanisms of<br />

mononuclear cells is important for the resolution of transplant destruction might be involved in glucoinflammation,<br />

which is coordinated by anti-inflam- corticoid efficacy.<br />

matory mediators derived from arachidonic acid<br />

(e.g. lipoxins and cyclopentane prostaglandins<br />

[cyPGs]) and by anti-inflammatory cytokines (e.g.<br />

4. Adverse Effects of Glucocorticoids<br />

IL-10, TGFβ). [149]<br />

It is a general observation that adverse effects of<br />

Glucocorticoids counteract the acute inflammaglucocorticoid<br />

treatment appear more likely after<br />

tory effects on the microcirculation resulting in<br />

long-term treatment but less frequently after shortvasoconstriction,<br />

reduction of oedema and a determ<br />

treatment, even with high glucocorticoid<br />

creased rate of leucocyte migration. [201] In addition,<br />

doses. [210] This observation is compatible with timeglucocorticoids<br />

could affect the resolution of independent<br />

genomic effects that do not increase furflammation<br />

by inducing annexin-1 derived peptides<br />

ther after high doses when glucocorticoid receptor<br />

that act at the lipoxin A4 receptor. [202] Glucocortisaturation<br />

is already achieved.<br />

coid treatment increases IL-10, [203] which inhibits<br />

NF-κB activity and affects many immune cells. [204] Glucocorticoids induce or aggravate diabetes<br />

mellitus and induce arterial hypertension. [211-213]<br />

3.4 Immune Response<br />

Glucocorticoids negatively regulate the hypothalamic-pituitary-adrenal<br />

(HPA) axis and rapidly inhibit<br />

The following events are involved in the immune adrenal secretion of cortisol. Long-term inhibition<br />

response: (i) antigen processing/presentation and acnal<br />

can lead to adrenal atrophy. [214] Unfortunately adre-<br />

tivation of macrophages/dendritic cells; (ii) antigen<br />

suppression cannot be predicted from the dose or<br />

recognition and activation of T cells; (iii) generation duration of glucocorticoid therapy. [215] One of the<br />

of T helper cell response; (iv) production of proinapy<br />

most serious adverse effects of glucocorticoid ther-<br />

flammatory cytokines; (v) adhesion and migration;<br />

is the increased risk of infection. [216,217] The risk<br />

(vi) inflammation and cell injury; and (vii) repair increases with the dose and duration of glucocortiand<br />

restitution. All of these events can be affected coid treatment.<br />

by glucocorticoids.<br />

Long-term treatment with glucocorticoids causes<br />

In acute kidney transplant rejection T helper cells osteoporosis by various mechanisms affecting osreact<br />

in a Th1 type immune response [205] after recogcally,<br />

teoblastic and osteoclastic functions. [211,218,219] Clininition<br />

of donor antigen. [206] Glucocorticoids act on<br />

the fracture risk increases with the glucocorti-<br />

various levels of transplant rejection. They inhibit coid dose [220] and correlates better with the daily<br />

the differentiation and antigen presentation of dose compared with the cumulative dose. [221] Skin<br />

macrophages and dendritic cells [184,207] and thus supferation<br />

atrophy is due to suppression of cutaneous cell proli-<br />

press the initiation of an immune response. Glucoglucocorticoids.<br />

and protein synthesis (e.g. collagen) by<br />

corticoids inhibit proinflammatory cytokine produccorticoid<br />

Skin thinning begins after glucocorticoids<br />

tion of IL-1, IL-2, IL-6, IL-12, IFNγ and TNFα in<br />

treatment for only a few days. [211]<br />

various cells, [208] leading to suppression of activated Other adverse effects associated with glucocorti-<br />

T cells. In addition, apoptosis of T cells is induced coid therapy include gastrointestinal ulcers (controby<br />

glucocorticoid pulse therapy, [190,191] which could versial), [211,222,223] cataract formation, [224] redistribualso<br />

be important in the treatment of acute transplant tion of body fat, dyslipidaemia, myopathy, bone<br />

rejection. Furthermore, down-regulation of adhesion necrosis, [225] growth retardation in children, [226]<br />

molecules and chemokine receptors that are up- glaucoma, psychosis, increased appetite and, rarely,<br />

© 2005 Adis Data Information BV. All rights reserved. Clin Pharmacokinet 2005; 44 (1)


Pharmacokinetics/Pharmacodynamics of Glucocorticoids 79<br />

allergic reactions. [227] Increased appetite can also be At the tissue/organ level, topical skin blanching<br />

a desired effect, e.g. in cancer patients. [228] as a result of vasoconstriction of the skin microvasculature<br />

has been used as a biomarker for dermato-<br />

5. Pharmacodynamics logical glucocorticoid products. [241,242]<br />

of Glucocorticoids<br />

The pharmacodynamic properties of a drug are<br />

5.2 Potency<br />

the mathematical description for the quantitative The effect (E), as measured by a change of a<br />

relationships between drug concentration and ef- biomarker, depends on drug concentration (C). The<br />

fects.<br />

basic correlation between effect and concentration is<br />

described by the sigmoid Emax model, where Emax is<br />

5.1 Biomarker and Surrogate Endpoints the maximum achievable effect (capacity parameter)<br />

and CE50 is the concentration producing the<br />

Many studies have analysed the pharmacohalf-maximal<br />

effect (sensitivity parameter). The latdynamics<br />

of glucocorticoids, but the results vary<br />

depending on the biomarker used. There are many<br />

ter has also been named EC50, CI50, or IC50, depen-<br />

biomarkers, but only a few well-evaluated surrogate<br />

ding on circumstances. The sigmoidicity constant H<br />

endpoints (e.g. HbA1c for complications of diabetes<br />

determines the slope of the curve (equation 1).<br />

H<br />

mellitus, bone mineral density for fracture risk in<br />

C<br />

E = E ·<br />

max H H<br />

osteoporosis).<br />

CE<br />

50<br />

+ C<br />

At the molecular level, endogenous cortisol is<br />

(Eq. 1)<br />

used as a biomarker for HPA suppression. [30] Osteo-<br />

The parameter CE50 is a hybrid of receptor affinicalcin,<br />

secreted by osteoblastic cells during ostety,<br />

the number of receptors, and subsequent effector<br />

ogenesis, is used as a biomarker for glucocorticoid<br />

mechanisms. The potency of a drug is calculated as<br />

effects on bone formation. [57,229] Molecular markers<br />

the inverse of CE50, since a high potency indicates<br />

used in vitro to evaluate glucocorticoid potencies<br />

that only a low concentration is needed to produce<br />

include transcription factors (e.g. AP-1, NF-κB),<br />

the half-maximal effect (equation 2).<br />

mRNA levels and gene expression (e.g. cytokines or<br />

membrane proteins). [230-234] 1<br />

Potency =<br />

At the cellular level, the number of T cells is a<br />

CE<br />

50<br />

frequently used biomarker in vivo because T cells<br />

(Eq. 2)<br />

play a central role in the immune response. [30,31,58-60] The relationship between drug concentrations<br />

The suppression of CD4+ (T helper) cells by gluco- and effects can be determined in vitro where concorticoids<br />

might be a surrogate endpoint for the stant concentrations and a defined exposure time are<br />

prevention and treatment of transplant rejection and used. Eventually an in vitro CE50 is calculated. In<br />

the therapy of autoimmune diseases. On the other vivo, drug concentrations are constantly changing<br />

hand, low CD4 counts in long-term immunosup- and the different half-lives of glucocorticoids influpression<br />

were correlated with an increased frequen- ence the duration of effect and obscure drug potency<br />

of infections, [235] neoplasms [236,237] and athero- cy. [243] The influence of both factors (CE 50 and drug<br />

sclerosis. [238] The number of circulating blood cells half-life) on the effect can be distinguished by pharreflects<br />

migration between the intravascular and ex- macokinetic/pharmacodynamic analysis and an estitravascular<br />

compartments, [31] but apoptosis might mation of in vivo CE50. Importantly, when in vitro<br />

play an additional role after very high glucocorticoid and in vivo CE 50 s are compared, protein binding has<br />

doses. [190,191] Lymphocyte activation and prolifera- to be considered. Such a comparison is further comtion<br />

are used as a biomarker in vitro. [61,239,240] plicated because local (biophase) concentrations<br />

© 2005 Adis Data Information BV. All rights reserved. Clin Pharmacokinet 2005; 44 (1)


80 Czock et al.<br />

have to be used depending on the selected biomarker.<br />

These local concentrations are determined by<br />

drug distribution and local glucocorticoid metabolism.<br />

contrast, using COX-2 expression as a biomarker,<br />

hydrocortisone had a lower CE50 (7.5 × 10 –8 mol/L)<br />

compared with dexamethasone (1 × 10 –7 mol/L).<br />

Furthermore, prednisolone inhibited COX-2 expression<br />

and not cPLA2 activity, but methylprednisolone<br />

inhibited cPLA2 activity and not COX-2 expres-<br />

sion. [160] These results can be explained only by<br />

distinct cellular pathways (i.e. nongenomic mechanisms<br />

in the case of cPLA2 inhibition). Another study<br />

analysed COX-2 expression and activity in human<br />

monocytes and found a low CE50 for dexameth-<br />

asone, a medium CE50 for methylprednisolone, and<br />

a high CE 50 for hydrocortisone, in agreement with<br />

receptor binding affinities. [232] These contradictory<br />

results could be due to the cell type studied (immune<br />

cell vs lung cell), but can also be explained by the<br />

design of the experiments. Croxtall et al. [160] used an<br />

incubation period of 3 hours, whereas Santini and<br />

colleagues [232] used a 24-hour period. The short in-<br />

cubation period of 3 hours is thought to favour<br />

nongenomic effects, whereas an incubation period<br />

of 24 hours would favour genomic effects.<br />

Nongenomic effects of glucocorticoids on cellular<br />

energy mechanisms have been suggested. [157]<br />

Cellular oxygen consumption is a biomarker for<br />

leucocyte activation, as immune functions require<br />

energy. Oxygen consumption by peripheral blood<br />

mononuclear cells was increased in patients with<br />

rheumatic disease and normalised after glucocorti-<br />

coid treatment. [248] Using cellular energy consump-<br />

tion as a biomarker, the following relationship was<br />

found in vitro: dexamethasone (1.2) > methylprednisolone<br />

(1.0) > prednisolone (0.4) [relative<br />

drug potencies compared with methylpred-<br />

The potency of a given glucocorticoid differs<br />

depending on biomarker and cell type, despite the<br />

glucocorticoid receptor being the same in different<br />

cells and tissues. Potency differences between bi-<br />

omarkers are explained by diverse effector mechan-<br />

isms. Potency differences between cell types might<br />

be explained by a different number of glucocorticoid<br />

receptors per cell, by a different glucocorticoid<br />

binding affinity due to the phosphorylation status of<br />

the glucocorticoid receptor, [143,144] possibly by glu-<br />

cocorticoid receptor diversity, [244] by intracellular<br />

modulation of the transactivation potency of the GR/<br />

GC complex by coactivators and corepressors, by a<br />

different histone deacetylase activity, [245] and possibly<br />

also by differences in nongenomic mechanisms<br />

between cell types.<br />

The potency for a given effect differs between the<br />

various glucocorticoids. This potency has been correlated<br />

with glucocorticoid affinity for the glucocor-<br />

ticoid receptor. [3,36,234,246] For example, there was a<br />

good correlation between the relative glucocorticoid<br />

receptor affinity of various glucocorticoids and the<br />

inhibition of whole-blood lymphocyte prolifera-<br />

tion. [247] How different glucocorticoid receptor affinities<br />

translate to different effects is not entirely<br />

clear. There is evidence that the molecular structure<br />

of different glucocorticoids does not influence the<br />

DNA binding affinity of the glucocorticoid receptor,<br />

but has allosteric effects on glucocorticoid receptor-<br />

DNA dissociation. [146] nisolone]. [161]<br />

The ranking of glucocorticoid potencies depends<br />

on the experimental design. Experiments looking at<br />

5.3 Clinical Efficacy<br />

genomic effects can lead to other rankings of<br />

glucocorticoid potency compared with nongenomic Clinical efficacy depends on pharmacodynamic<br />

effects. For example, using cPLA2 activation as a (e.g. potency) and pharmacokinetic (e.g. duration of<br />

biomarker, dexamethasone had a lower CE50 (2 × the drug at the receptor site) characteristics of a<br />

10–8 mol/L), and thus higher potency compared with drug. Taken together, both parameters determine the<br />

hydrocortisone (CE50 7.5 × 10 –8 mol/L) in A549 duration of the momentary effect, which in turn<br />

cells (a human lung adenocarcinoma cell line). In correlates with overall clinical efficacy:<br />

© 2005 Adis Data Information BV. All rights reserved. Clin Pharmacokinet 2005; 44 (1)


Pharmacokinetics/Pharmacodynamics of Glucocorticoids 81<br />

potency + presence at the receptor site<br />

→ momentary effect course (effect duration)<br />

sum of momentary effects → clinical efficacy<br />

Mathematically, pharmacokinetic/pharmacodynamic<br />

models describe the momentary effect course.<br />

This effect course is characterised by the duration of<br />

effect (i.e. the time until the effect drops below a<br />

specified value). The momentary effect course can<br />

be summed up as the area under the effect-time<br />

curve (AUETC or AUEC) or as the area between the<br />

baseline and effect curve (ABEC), which are assumed<br />

to correlate with the clinical effect. [249]<br />

5.4 Pharmacodynamic<br />

Drug-Drug Interactions<br />

Drug-drug interactions can occur on a pharmacodynamic<br />

level when the same effector pathways<br />

are involved. For example, synergistic effects<br />

of prednisolone, ciclosporin and sirolimus on lymphocyte<br />

proliferation have been demonstrated in<br />

vitro. [250] Recombinant human IL-10 and prednisolone<br />

additively inhibit lymphocyte proliferation in<br />

vitro. [239] In vivo, IL-10 increases prednisolone-induced<br />

lymphocyte suppression and neutrophil stimulation,<br />

but decreases prednisolone-induced monocyte<br />

suppression. These changes were due to altered<br />

glucocorticoid potency whereas pharmacokinetics<br />

were unchanged. [62]<br />

Theophylline has a glucocorticoid-sparing effect<br />

in patients with asthma, [251] which can be explained<br />

on a molecular level by enhanced HDAC activity in<br />

epithelial cells and macrophages. [252] In contrast to<br />

theophylline, smoking reduces HDAC activity,<br />

which might explain the enhanced expression of<br />

inflammatory mediators in smokers. [253] Clarithromycin<br />

increases the glucocorticoid sensitivity of<br />

lymphocytes from patients with asthma as measured<br />

by suppression of lymphocyte activation in vitro. [254]<br />

Glucocorticoids and β 2 -agonists act synergistically<br />

on bronchial smooth muscle cell proliferation as<br />

shown in vitro. [255]<br />

6. Pharmacokinetic/<br />

Pharmacodynamic Models<br />

Generally, mathematical models can be put into<br />

two categories: (i) models of data, also called empirical<br />

or descriptive models; and (ii) models of sys-<br />

tem, also called mechanistic or explanatory models.<br />

[256,257] For predictions, usually a mechanistic<br />

model and a critical evaluation of the model is<br />

required. Predictions can include the pharmacody-<br />

namic response to altered dosage regimens (i.e.<br />

dose, timing, route of administration), [63,258,259] the<br />

pharmacodynamic response to altered pharmaco-<br />

kinetics (e.g. elimination conditions), or prediction<br />

of clinical pharmacodynamics based on in vitro<br />

data. [3,5]<br />

The pharmacokinetic part of the pharmacokinetic/pharmacodynamic<br />

model depends on the pharmacokinetics<br />

of the administered drug. Free drug<br />

concentrations should be measured if the pharmaco-<br />

kinetics of total drug concentrations are nonlinear or<br />

if inclusion of constants from in vitro measurements<br />

in the pharmacodynamic model is desired. The phar-<br />

macodynamic part of the pharmacokinetic/pharma-<br />

codynamic model depends on the selected bi-<br />

omarker. In the case of receptor-mediated effects,<br />

the Emax and the sigmoid Emax models are the most<br />

widely used. [4] Generally, the simplest model, which<br />

explains observations satisfactorily and makes cor-<br />

rect predictions of future experiments, is regarded as<br />

an appropriate model. Therefore the sigmoid Emax<br />

model (using three model parameters) should only<br />

be chosen if the simple Emax model (using two<br />

model parameters) is not sufficient. Pharmacokine-<br />

tic/pharmacodynamic models including clinical<br />

endpoints (e.g. transplant rejection, response rate,<br />

disease progression [260] ) have not been published yet<br />

for glucocorticoids.<br />

A lag between the time of maximum effect and<br />

the time of maximum concentration may be ex-<br />

plained by the link in the pharmacokinetic and pharmacodynamic<br />

model. Firstly, a hypothetical effect<br />

compartment can be presumed (biophase distribu-<br />

tion model). [1,4] However, it has been argued that<br />

© 2005 Adis Data Information BV. All rights reserved. Clin Pharmacokinet 2005; 44 (1)


82 Czock et al.<br />

this approach is only valid if the lag time is due to pools by physiological mechanisms needs some<br />

protracted drug delivery to the biophase. Secondly, time, and the effect delay is thus explained by a<br />

an indirect response model should be used if the physiological mechanism. [30,31,58,61,84] A comparison<br />

mechanism is, for example, inhibition of the produc- between an indirect response model (formerly<br />

tion of a biomarker, such that the observed decrease named direct suppression model) and an effect comin<br />

this biomarker depends on time-limited elimina- partment model (using the Emax model, the sigmoid<br />

tion through other mechanisms. [1,261] Both models E max model or a threshold E max model) for gluco-<br />

(effect compartment vs indirect response model) can corticoid effects (biomarker: basophil count) remimic<br />

each other under certain circumstances, [262] vealed that the indirect response model provided<br />

but model selection based on mechanistic know- more consistent results. [6]<br />

ledge of the drug action should provide a more Inclusion of a baseline function into the model is<br />

reliable model. [263] Thirdly, a transduction model necessary for biomarkers that display a circadian<br />

should be used if time-dependent transduction pro- pattern (e.g. endogenous cortisol, blood lymphocesses<br />

are involved. [1,61,264] A fourth option is to cytes, osteocalcin [265] ). A number of methods to<br />

include a lag time. [3,64,65] The latter option lacks a account for baseline variation of endogenous corphysiological<br />

basis, but satisfactory predictions tisol have been compared recently and the method<br />

have also been possible with such a model. [3,65] using Fourier analysis proved to be the best. [266]<br />

Simpler and more often used methods are the use of<br />

6.1 Pharmacokinetic/Pharmacodynamic sinus functions [24,30,61] and the use of a linear release<br />

Analysis of Glucocorticoids<br />

model. [267,268] When baseline variation is disregarded,<br />

the effect might be overestimated and the con-<br />

Effect compartment models have been used for centration CE50 would be lower compared with a<br />

the description of lymphocyte suppression by model where a baseline function is included (e.g.<br />

glucocorticoids, which is maximal 4–6 hours after CE 50 = 1.5 µg/L [64] vs 20.12 µg/L [68] for T helper<br />

the maximum glucocorticoid concentration is cell suppression by methylprednisolone). Advanced<br />

reached. [60,66,67] However, it is unlikely that drug models can also explain the circadian pattern of<br />

distribution to the biophase is the physiological lymphocytes by the circadian pattern of cortisol<br />

mechanism that explains the delay of the glucocorti- secretion. [31,59]<br />

coid effect on lymphocytes. Therefore, using lym- Sophisticated mechanistic models including mophocyte<br />

suppression as a biomarker, the effect com- lecular aspects like glucocorticoid receptor downpartment<br />

model would be classified as a descriptive regulation have been developed and evaluated in<br />

model. Notably, the estimated CE 50 values varied animal models. [133,269]<br />

depending on the applied dose level, [60,67] which<br />

indicates selection of the wrong model because<br />

CE50, as a parameter of the drug, should be independent<br />

6.2 In Vivo Potency<br />

of the dosage scheme. However, another expla- We have summarised the in vivo CE50 values of<br />

nation could be the measurement of total and not selected glucocorticoids in table V. In order to comfree<br />

plasma concentrations of prednisolone in these pare different glucocorticoids, a potency ratio can be<br />

studies.<br />

calculated from their CE 50 values for a given bi-<br />

Indirect response models have been used for omarker. For example, a potency ratio of 17.1 was<br />

glucocorticoid effects on blood cell count and en- found for T helper cell suppression by methyldogenous<br />

cortisol. In these models, cell migration or prednisolone compared with cortisol. [31] Comparing<br />

secretion of cortisol into the blood is blocked by the CE50 values for various biomarkers of estimated<br />

glucocorticoids. Elimination of pre-existing blood free methylprednisolone and free prednisolone<br />

© 2005 Adis Data Information BV. All rights reserved. Clin Pharmacokinet 2005; 44 (1)


© 2005 Adis Data Information BV. All rights reserved. Clin Pharmacokinet 2005; 44 (1)<br />

Table V. Glucocorticoid pharmacodynamics. The concentrations producing the half-maximum effect CE50 (µg/L) of selected glucocorticoid effects are shown. The pooled mean ±<br />

pooled standard deviation (see Appendix) and the range of the primary mean values (minimum–maximum) are given for cases where more than one study was<br />

found [3,6,24,25,28-31,36,40,43,55,57-59,61-68]<br />

Drug Drug plasma Cellular effects Adverse effects<br />

concentration lymphocyte T helper cell cytotoxic T basophil monocyte neutrophil cortisol glucose osteocalcin water<br />

measured suppression suppression cell a suppression suppression induction suppression induction suppression retention<br />

suppression<br />

Cortisol/ Total 6.7 ± 4.9 b 67.9 ± 20.7 103.0 ± 17.5 179.0 ± 66.2 4.6 ± 5.3<br />

hydrocortisone (56.4–79.3)<br />

Free 15.4 ± 3.4<br />

Prednisolone Total 125.3 ± 65.2 61.9 ± 40.6 9.7 ± 3.4<br />

(90.4–173.9) (48.8–75.0) (9.0–10.3)<br />

Free 9.7 ± 7.8 5.8 ± 6.7 12.1 ± 8.1 5.7 ± 6.4 10.5 ± 6.1 29.2 ± 31.5 0.9 ± 1.3 4.5<br />

(4.7–15.1) (3.2–15.1) (4.8–19.2) (15.0–57.7) (0.3–1.9)<br />

Methylprednisolone Total c 10.5 ± 11.8 ± 11.5 d 51.0 ± 43.3 7.2 ± 6.1 22.4 ± 27.5 1.0 ± 1.1 102.7 8.8<br />

13.9 (1.4–22.8) (18.5–112.0) (2.1–14.3) (8.4–36.0) (0.1–2.9)<br />

(6.0–13.9)<br />

Dexamethasone Total c 2.9 ± 1.0 4.0 ± 4.7 18.0 ± 4.7 3.7 ± 5.3 0.1 ± 0.07 26.9<br />

a<br />

(0.9–6.0) (2.2–6.1)<br />

These CD3+/CD8+ cells were named T suppressor cells in other studies.<br />

b This very low value is most probably due to the lack of a baseline function in the original model. [66]<br />

c<br />

When free concentrations were given, these were converted to total concentrations using the free fraction given in the respective study.<br />

d One study used the sigmoid Emax model and estimated the sigmoidicity parameter H as 1.2 ± 0.1. [64]<br />

Emax = maximal effect; Free = unbound to plasma components; Total = plasma bound and free.<br />

Pharmacokinetics/Pharmacodynamics of Glucocorticoids 83


84 Czock et al.<br />

(table V), we find potency ratios between 1 and 6. In 6.3.3 Once- Versus Twice-Daily<br />

Glucocorticoid Administration<br />

accordance with these ratios, a retrospective study<br />

of kidney transplant patients suggested that methylolone<br />

and methylprednisolone concentrations fall to<br />

Due to their short elimination half-lives, prednis-<br />

prednisolone might be superior to prednisolone for<br />

below their CE50 values (biomarker: T helper cells)<br />

maintenance immunosuppression. [270]<br />

within 12 hours after administration (low dose).<br />

Eventually the number of T helper cells increases<br />

6.3 Selected Examples again and also shows a small rebound. This rebound<br />

can be explained by ongoing suppression of endogenous<br />

cortisol (the CE50 for cortisol suppression is<br />

6.3.1 Influence of Sex on Glucocorticoid<br />

lower than the CE50 for T helper cell suppression),<br />

Pharmacokinetic/Pharmacodynamic Properties<br />

which in turn leads to further diminished T helper<br />

Pharmacokinetic differences between males and<br />

cell suppression.<br />

females have been reported for prednisolone and<br />

As T helper cells play a central role in transplant<br />

methylprednisolone. The prednisolone clearance is<br />

rejection, it has been suggested that the efficacy of<br />

lower [84] but methylprednisolone clearance is glucocorticoids might be increased by twice-daily<br />

higher [68] in females compared with males. How- compared with the traditional once-daily adminisever,<br />

based on pharmacodynamic measurements, it tration. [64,67] Thus, glucocorticoid efficacy would be<br />

was concluded that dose adjustment is not neces- greater, which might allow a lower total daily<br />

sary. [68,84] For example, the higher clearance of dose. [63] Theoretically, adverse effects with CE50<br />

methylprednisolone in women was compensated by values higher or equal compared to the peak concentration<br />

a higher potency (i.e. a lower CE 50 ), leading to the<br />

should be reduced in parallel with the total<br />

same overall effect. [68] The conclusions from these daily dose.<br />

studies are limited to the biomarkers employed, but In order to compare once-daily with twice-daily<br />

they underline the importance of performing pharmethylprednisolone<br />

administration we performed a simulation with<br />

macodynamic analyses.<br />

8mg once-daily versus 4mg<br />

twice-daily. We used an established model and published<br />

6.3.2 Influence of Age on Glucocorticoid<br />

parameter values for methylprednisolone ef-<br />

Pharmacokinetic/Pharmacodynamic Properties fects on T helper cells and endogenous corti-<br />

The frequency and severity of adverse events<br />

sol. [30,273] The pharmacokinetic part of the model<br />

may be increased in elderly patients. Baseline cornamic<br />

part was a precursor-dependent indirect res-<br />

was a one-compartment model and the pharmacodytisol<br />

is similar in young and elderly males, but<br />

ponse model (see Appendix). Simulations were<br />

adrenal suppression (biomarker: endogenous cordone<br />

with the use of the software WinNonlin Profestisol)<br />

after administration of methylprednisolone<br />

sional 4.0.1 (Pharsight Corporation, California).<br />

was greater in the elderly. [271] This can be explained<br />

The total effect correlates with the area between the<br />

by a decrease in methylprednisolone clearance to<br />

baseline curve without drug administration and the<br />

66% and an increase in the half-life to 130% in the effect curve after drug administration. This simulaelderly<br />

(69–82 years) males. [69] Also prednisolone tion showed that twice-daily administration has a<br />

clearance was lower (62%) in elderly subjects greater total effect despite the same total dose (fig-<br />

(65–89 years). [272] However, in the latter study adre- ure 3). The total effect of methylprednisolone on T<br />

nal suppression (biomarker: endogenous cortisol) helper cells is stronger and the rebound of T helper<br />

after administration of prednisolone was lower in cells is lower after twice daily administration. Thus,<br />

the elderly.<br />

a lower total dose might be possible. However,<br />

© 2005 Adis Data Information BV. All rights reserved. Clin Pharmacokinet 2005; 44 (1)


Pharmacokinetics/Pharmacodynamics of Glucocorticoids 85<br />

T helper cells (cells/µL)<br />

1250<br />

1000<br />

750<br />

500<br />

250<br />

T helper cells without methyprednisolone<br />

administration (baseline curve)<br />

Methylprednisolone 8mg once daily<br />

Methylprednisolone 4mg twice daily<br />

0<br />

0<br />

8:00 16:00 0:00 8:00<br />

200<br />

150<br />

100<br />

Time (h)<br />

Fig. 3. Effect of methylprednisolone on T helper cells after 8mg<br />

once or 4mg twice daily administration. Methylprednisolone concentrations<br />

(lower two curves) and effect on T helper cell counts<br />

(upper three curves) after methylprednisolone were simulated using<br />

an established model and published parameter values (see Appendix).<br />

[30]<br />

adverse effects with low CE50 values, such as sup-<br />

pression of endogenous cortisol, might also be<br />

greater (figure 4). This simulation is in agreement<br />

with measured concentrations and effects after fractionated<br />

dose administration. Administration of<br />

twice-daily dose fractions increased the immunosuppressive<br />

efficacy of methylprednisolone (biomarker:<br />

T helper cell suppression) while it did not<br />

increase short-term adverse effects (biomarker: insulin<br />

and glucose), except for endogenous cortisol.<br />

[64]<br />

6.3.4 Adverse Effects<br />

The CE50 for glucose induction is higher than the 100<br />

CE50 for lymphocyte suppression. [3,65] Thus, if a<br />

smaller daily dose is possible, e.g. with administration<br />

of twice-daily doses, the overall effect of gluco-<br />

50<br />

corticoids on glucose metabolism should be less.<br />

The degree and duration of endogenous cortisol<br />

0<br />

suppression is dose dependent. [7] However, it is unclear<br />

whether the degree of short-term cortisol suppression<br />

correlates with adrenal atrophy and limited<br />

adrenal reserve or not. [215]<br />

50<br />

Methylprednisolone (µg/L)<br />

7. Clinical Aspects of<br />

Glucocorticoid Therapy<br />

Clinically, glucocorticoid dosage practice can be<br />

divided into five categories. These are low dose<br />

(≤7.5mg prednisolone equivalent per day), medium<br />

dose (>7.5mg, but ≤30mg prednisolone equivalent<br />

per day), high dose (>30mg, but ≤100mg prednisolone<br />

equivalent per day), very high dose (>100mg<br />

prednisolone equivalent per day), and pulse therapy<br />

(≥250mg prednisolone equivalent per day for one or<br />

a few days). The first three categories were estimated<br />

to correspond to


86 Czock et al.<br />

macodynamics (e.g. includes additional nongenom- methylprednisolone pulse therapy. [283,284] In a retroic<br />

mechanisms) compared with maintenance immu- spective study of diffuse alveolar haemorrhage assonosuppression.<br />

Clinically, the effect of glucocorti- ciated with bone marrow transplantation, only very<br />

coid pulses is more similar to an all-or-none high doses of glucocorticoids were effective. [285]<br />

principle than to a steady function of dose. Some<br />

Treatment with dexamethasone has been sugdiseases<br />

(other than those requiring emergency thergested<br />

for acute CNS diseases. [286,287] Dexamethapy)<br />

are treated intermittently with glucocorticoid<br />

asone is usually preferred for treatment of CNS<br />

pulses in order to maximise the beneficial effects as<br />

diseases, because penetration into cerebrospinal fluwell<br />

as to minimise adverse effects.<br />

id is superior compared with prednisolone, as shown<br />

Glucocorticoid pulses are usually administered<br />

in a nonhuman primate model. [288]<br />

intravenously. However, very high oral doses (≥1g)<br />

of prednisolone (as tablets) and methylprednisolone<br />

7.1.2 Non-Emergency Treatment<br />

succinate (as an oral solution) had maximum concentrations<br />

and area under the plasma concentramanagement<br />

Glucocorticoids are a standard therapy for the<br />

of acute SLE. [289,290] Despite an in-<br />

tion-time curves (AUCs) similar to those after<br />

intravenous administration. [15,275,276] Therefore, oral<br />

creased risk of infections with glucocorticoid treatadministration<br />

of glucocorticoids might be investiprednisolone<br />

ment, overall survival is improved. [291] Methyl-<br />

gated for some indications.<br />

pulses are effective [292] and there was a<br />

more rapid improvement in renal function following<br />

7.1.1 Emergency Treatment<br />

pulse methylprednisolone therapy in lupus nephritis<br />

[293,294] while adverse effects were not in-<br />

Antirejection therapy with intravenous glucocorcreased.<br />

In patients with nonrenal lupus<br />

ticoid 1000mg pulses was first reported in patients<br />

erythematosus, life-threatening manifestations such<br />

with kidney transplants. [277] The clinical response of<br />

as coma, seizures or thrombocytopenia responded<br />

antirejection treatment correlated with the induction<br />

better to methylprednisolone pulses than to oral<br />

of apoptosis of infiltrating lymphocytes. [278] This<br />

glucocorticoid protocols.<br />

mechanism might contribute to the fast clinical ef-<br />

[296]<br />

fect on the patient’s symptoms and objective paragroup<br />

Primary glomerulonephritis is a heterogeneous<br />

meters such as transplant size after such pulses.<br />

of diseases that can lead to the nephrotic<br />

Crescentic rapidly progressive glomerulonephritis<br />

syndrome, chronic renal failure or both. IgA nephritis,<br />

mainly the type not due to antiglomerular basement<br />

can be treated with methylprednisolone pulses<br />

membrane antibodies, was treated successfully (1000mg for 3 days) every 2 months. The mainten-<br />

with methylprednisolone intravenous pulse therapy ance dose is administered orally on alternate<br />

(1000 mg/day for 3–5 days [279] or 7 days [280] ). days. [297] Previous studies that did not use glucocor-<br />

Atheroembolic renal disease, caused by cholestic<br />

ticoid pulses were unable to demonstrate a therapeuterol<br />

emboli, is associated with eosinophilia and<br />

effect of glucocorticoids in IgA nephritis. In<br />

vasculitis-like histological appearance [281] and was severe idiopathic childhood nephrotic syndrome,<br />

effectively treated by methylprednisolone pulses methylprednisolone 1g/1.73m 2 pulses induced a<br />

(500mg for 3 days), followed by oral prednisone 0.5 more rapid remission of proteinuria than oral predmg/kg/day.<br />

[282]<br />

nisolone. [298]<br />

Patients with diffuse alveolar haemorrhage due to Glucocorticoids are a standard therapy for the<br />

SLE, systemic vasculitis or anti-GBM (glomerular management of multiple myeloma. Very high-dose<br />

basement membrane) nephritis such as Goodpasrelapsed<br />

methylprednisolone has been used in refractory or<br />

ture’s syndrome were treated effectively with<br />

[299]<br />

multiple myeloma.<br />

© 2005 Adis Data Information BV. All rights reserved. Clin Pharmacokinet 2005; 44 (1)


Pharmacokinetics/Pharmacodynamics of Glucocorticoids 87<br />

7.2 Diseases Not Treated with Pulse Therapy twice-daily administration in glucocorticoid replacement<br />

therapy. [313]<br />

Rheumatoid arthritis is characterised by the acsystemic,<br />

Maintenance therapy of asthma is based not on<br />

cumulation and persistence of inflammatory cells in<br />

but on local (inhaled) glucocorticoid ad-<br />

synovial joints, which results in joint damage. The ministration. [242,246] A more frequent administration<br />

anti-inflammatory and antiproliferative effects of of budesonide (four times daily vs twice daily) was<br />

glucocorticoids [300] leads to a reduced rate of joint superior in patients with moderate to severe asth-<br />

destruction. [301]<br />

ma. [314]<br />

Systemic glucocorticoids are used in the treatment<br />

of acute asthma. Oral treatment was similarly 7.4 Variation in Clinical Response to<br />

effective compared with intravenous methylprednisolone<br />

Glucocorticoid Therapy<br />

pulses. [302] Very high doses of hydro-<br />

cortisone provided no benefit compared with lower The clinical response to glucocorticoid treatment<br />

doses. [303] Methylprednisolone is often preferred for varies considerably between patients. This variabiltreatment<br />

of lung diseases because it achieves higher ity can be due to pharmacokinetic properties, pharconcentrations<br />

in the lung compared with prednisoticoid<br />

clearance could lead to lower exposure to the<br />

macodynamic properties, or both. Higher glucocor-<br />

lone (as measured in bronchoalveolar lavage fluid of<br />

rabbits). [304]<br />

drug and consequently to lower immunosuppressive<br />

Severe sepsis and the adult respiratory distress<br />

and anti-inflammatory action. For example, patients<br />

syndrome involve an uncontrolled host defence resrejection<br />

had a shorter methylprednisolone half-<br />

with kidney transplants who experienced transplant<br />

ponse that includes inflammation, endothelial damlife.<br />

age, enhanced coagulation, microthrombi and relative<br />

[315]<br />

adrenal insufficiency. [305] Treatment with very A correlation between the clinical response and<br />

high bolus doses of methylprednisolone for 24 hours the in vitro responsiveness of stimulated lympho-<br />

did not improve mortality, [306,307] but prolonged cytes to glucocorticoid treatment has been observed.<br />

treatment with stress-dose glucocorticoids for days Such a correlation might be useful to predict the<br />

to weeks was beneficial. [305,308,309]<br />

clinical response in patients with focal and segmen-<br />

tal sclerosing glomerulonephritis, [316] with kidney<br />

transplants, [317,318] with asthma, [319] and with rheu-<br />

7.3 Role of the Dosage Interval matoid arthritis. [320]<br />

8. Conclusions<br />

At one time, glucocorticoids were thought to be<br />

qualitatively indistinguishable [321] because they act<br />

via the same receptor, but today qualitative differ-<br />

ences have been discovered. Many beneficial and<br />

adverse effects of glucocorticoids are due to geno-<br />

mic mechanisms, but there is growing evidence that<br />

some glucocorticoid effects are mediated by nonge-<br />

nomic mechanisms, especially with pulse glucocor-<br />

ticoid therapy. As these mechanisms can differ between<br />

the various glucocorticoids, one glucocorti-<br />

coid cannot be simply replaced by another.<br />

A twice-daily glucocorticoid dosage is recommended<br />

clinically for treatment of severe diseases.<br />

[310] Daily prednisolone produced more intensive<br />

and longer sustained immunosuppressive effects<br />

(biomarker: T cells) than alternate-day<br />

treatment in kidney transplant patients. [311] Accordingly,<br />

prolongation of the dosage interval should<br />

lead to reduced effects. A prednisolone dosage regimen<br />

of 90mg on alternate days was less effective<br />

than 15mg every 8 hours in patients with giant cell<br />

arteritis. [312]<br />

Due to the short half-life of hydrocortisone, administration<br />

three times daily might be superior to<br />

© 2005 Adis Data Information BV. All rights reserved. Clin Pharmacokinet 2005; 44 (1)


88 Czock et al.<br />

Pharmacokinetic/pharmacodynamic models are<br />

useful for simultaneously analysing the pharmacokinetics<br />

and pharmacodynamics of a drug and for<br />

making predictions. For low-dose maintenance therapy<br />

with glucocorticoids, twice-daily dose fractions<br />

might allow a lower daily dose and possibly a reduction<br />

in some adverse effects. Intravenous glucocorticoid<br />

pulses can be given for some indications with a<br />

dose interval of 4 weeks. New pharmacokinetic/<br />

pharmacodynamic models for glucocorticoid pulse<br />

therapy should be developed and evaluated.<br />

Acknowledgements<br />

This study was supported by the European Commission<br />

within the PharmDIS project (BMH4-CT98-9548 and IST<br />

Craft-2001-52107). The authors have no conflicts of interest<br />

to disclose.<br />

Appendix<br />

Statistical Data Synthesis<br />

Published pharmacokinetic parameters are heter-<br />

ogeneous and their values vary between studies.<br />

Therefore, a statistical data synthesis is necessary to<br />

combine such values and estimate the population<br />

mean. [322] In the current paper we summarised the<br />

published values from different publications as the<br />

pooled mean x where x , x , ...<br />

1 2<br />

xk<br />

are the published<br />

mean values and the total number of subjects n was<br />

calculated as n = n 1 + n2 +…+ nk (equation 3).<br />

Model for Simulation<br />

We used an indirect response model for simula-<br />

tion of twice-daily dose fractions of methylprednisolone<br />

(figure 3 and figure 4). The pharmaco-<br />

kinetic model was a one-compartment model with a<br />

first-order formation rate and a first-order elimination<br />

rate where MPs is methylprednisolone succi-<br />

nate, MP is methylprednisolone, k f is the formation<br />

rate constant, and ke is the elimination rate constant<br />

(equation 5 and equation 6).<br />

dMPs<br />

kf MPs<br />

dt<br />

= - ·<br />

dMP<br />

dt<br />

k<br />

= ·<br />

f<br />

MPs<br />

k<br />

- ·<br />

e<br />

MP<br />

(Eq. 5)<br />

(Eq. 6)<br />

The pharmacodynamic model for T helper cell<br />

suppression was a precursor-dependent indirect response<br />

model as developed by Sharma et al. [273] and<br />

Booker et al. [30] ThE are extravascular T helper cells<br />

and Th are blood T helper cells. The rate constant kin<br />

describes formation of new T helper cells, the timevarying<br />

rate constant kp(t) describes the migration of<br />

T helper cells from the extravascular to the blood<br />

compartment, the rate constant kout describes the<br />

removal of T helper cells, and I(t) describes the<br />

inhibitory effect of glucocorticoids on cell migration<br />

(equation 7 and equation 8).<br />

dThE<br />

= kin<br />

- kp<br />

( t)<br />

· I(<br />

t)<br />

· ThE<br />

dt<br />

dTh<br />

= kp<br />

( t)<br />

I(<br />

t)<br />

dt<br />

· ·<br />

ThE<br />

-<br />

k<br />

·<br />

out<br />

Th<br />

(Eq. 7)<br />

n1<br />

x n x n x<br />

x<br />

· 1<br />

+ 2<br />

· 2<br />

+ … + k · k<br />

=<br />

(Eq. 8)<br />

n<br />

The inhibitory function I(t) for the effect of<br />

(Eq. 3) methylprednisolone was an E max model where para-<br />

The pooled standard deviation was calculated meter Emax is the maximum achievable effect and<br />

using the published standard deviations s1, s2, …sk parameter CE50 is the concentration producing the<br />

(equation 4). half-maximal effect (equation 9).<br />

2<br />

2<br />

2<br />

( ) ( ) ( ) ½<br />

é s1 n1<br />

−1<br />

+ s2<br />

n2<br />

−1<br />

+ … + sk<br />

nk<br />

-1<br />

ù<br />

s =<br />

ê<br />

ú<br />

ë<br />

n - k<br />

û<br />

I(<br />

t)<br />

=<br />

1<br />

-<br />

E<br />

CE<br />

·<br />

max<br />

50<br />

+<br />

MP(<br />

t)<br />

MP(<br />

t)<br />

(Eq. 4) (Eq. 9)<br />

© 2005 Adis Data Information BV. All rights reserved. Clin Pharmacokinet 2005; 44 (1)


Pharmacokinetics/Pharmacodynamics of Glucocorticoids 89<br />

For the migration of T helper cells from the pharmacodynamic parameters for cortisol suppresextravascular<br />

to the blood compartment a periodic sion: CE50 = 0.446 µg • L –1 , Emax = 1; system<br />

function was applied where Rm is the mean input parameters of cortisol variation: kout = 0.338 h –1 ,<br />

rate, Rb is the amplitude and tz is the peak time in tmin = 12.2 hours (24-hour clock), tmax = 21.4<br />

relation to time zero (equation 10).<br />

hours (24-hour clock), Rm = 23.5 µg • L –1 • h –1 , Rb<br />

é 2π<br />

ù<br />

= 22.5 µg • L –1 • h –1 .<br />

k = + · ( - ) ·<br />

p<br />

( t)<br />

Rm<br />

Rb<br />

cos t t<br />

ê z ú<br />

24 û<br />

ë<br />

k ( t)<br />

in<br />

k ( t)<br />

in<br />

=<br />

=<br />

R<br />

R<br />

m<br />

+<br />

t = T min to T max<br />

=<br />

m<br />

+<br />

R<br />

m<br />

+<br />

R<br />

·<br />

b<br />

·<br />

b<br />

t = T max to 24 hours<br />

k ( t)<br />

in<br />

R<br />

R<br />

é 2π<br />

( t<br />

cos<br />

2<br />

ê<br />

ë<br />

ê<br />

ë<br />

é2π<br />

cos<br />

·<br />

b<br />

cos<br />

é<br />

ê<br />

·<br />

ë2<br />

24<br />

T )<br />

24)<br />

· + -<br />

max<br />

·(<br />

Tmin<br />

- Tmax<br />

+<br />

( t<br />

2<br />

T<br />

T<br />

ù<br />

ú<br />

û<br />

(Eq. 12)<br />

· - ·<br />

min<br />

-<br />

max<br />

ù<br />

ú<br />

2 · ( Tmax<br />

- Tmin<br />

)<br />

û<br />

2π<br />

( t<br />

( T<br />

T<br />

· -<br />

max<br />

min<br />

- Tmax<br />

+<br />

)<br />

)<br />

24)<br />

(Eq. 13)<br />

ù<br />

ú<br />

û<br />

(Eq. 10) References<br />

The pharmacodynamic model for cortisol sup-<br />

pression used the same inhibitory function I(t). The 510-8<br />

time-varying secretion of cortisol was described by<br />

kin(t) and the elimination of cortisol by the rate<br />

18-31<br />

constant kout (equation 11).<br />

dCort<br />

= kin<br />

( t)<br />

· I(<br />

t)<br />

- k ·<br />

out<br />

Cort<br />

dt<br />

(Eq. 11)<br />

The input function kin(t) for secretion of cortisol<br />

was a dual cosine model that also allows for asymmetric<br />

inputs where Tmax and Tmin are the timepoints<br />

where the secretion is maximal and minimal, respectively<br />

(equations 12, 13 and 14).<br />

t = 0 to T min<br />

1. Mager DE, Wyska E, Jusko WJ. Diversity of mechanism-based<br />

pharmacodynamic models. Drug Metab Dispos 2003; 31 (5):<br />

2. Meibohm B, Derendorf H. Pharmacokinetic/pharmacodynamic<br />

studies in drug product development. J Pharm Sci 2002; 91 (1):<br />

3. Derendorf H, Hochhaus G, Mollmann H, et al. Receptor-based<br />

pharmacokinetic-pharmacodynamic analysis of corticosteroids.<br />

J Clin Pharmacol 1993; 33 (2): 115-23<br />

4. Holford NH, Sheiner LB. Understanding the dose-effect relationship:<br />

clinical application of pharmacokinetic-pharmacodynamic<br />

models. Clin Pharmacokinet 1981; 6 (6): 429-53<br />

5. Mager DE, Jusko WJ. Quantitative structure-pharmacokinetic/<br />

pharmacodynamic relationships of corticosteroids in man. J<br />

Pharm Sci 2002; 91 (11): 2441-51<br />

6. Kong AN, Ludwig EA, Slaughter RL, et al. Pharmacokinetics<br />

and pharmacodynamic modeling of direct suppression effects<br />

of methylprednisolone on serum cortisol and blood histamine<br />

in human subjects. Clin Pharmacol Ther 1989; 46 (6): 616-28<br />

7. Mollmann H, Rohdewald P, Barth J, et al. Pharmacokinetics<br />

and dose linearity testing of methylprednisolone phosphate.<br />

Biopharm Drug Dispos 1989; 10 (5): 453-64<br />

8. Mollmann H, Rohdewald P, Barth J, et al. Comparative<br />

pharmacokinetics of methylprednisolone phosphate and<br />

hemisuccinate in high doses. Pharm Res 1988; 5 (8): 509-13<br />

9. Derendorf H, Mollmann H, Barth J, et al. Pharmacokinetics and<br />

oral bioavailability of hydrocortisone. J Clin Pharmacol 1991;<br />

31 (5): 473-6<br />

10. Kawai S, Ichikawa Y, Homma M. Differences in metabolic<br />

properties among cortisol, prednisolone, and dexamethasone<br />

in liver and renal diseases: accelerated metabolism of dexamethasone<br />

in renal failure. J Clin Endocrinol Metab 1985; 60<br />

(5): 848-54<br />

11. Hare LE, Yeh KC, Ditzler CA, et al. Bioavailability of dexamethasone.<br />

II: dexamethasone phosphate. Clin Pharmacol<br />

Ther 1975; 18 (3): 330-7<br />

12. Varis T, Kivisto KT, Backman JT, et al. The cytochrome P450<br />

3A4 inhibitor itraconazole markedly increases the plasma concentrations<br />

of dexamethasone and enhances its adrenal-suppressant<br />

effect. Clin Pharmacol Ther 2000; 68 (5): 487-94<br />

13. Toth GG, Kloosterman C, Uges DR, et al. Pharmacokinetics of<br />

high-dose oral and intravenous dexamethasone. Ther Drug<br />

Monit 1999; 21 (5): 532-5<br />

14. Garg V, Jusko WJ. Bioavailability and reversible metabolism of<br />

prednisone and prednisolone in man. Biopharm Drug Dispos<br />

1994; 15 (2): 163-72<br />

15. Patel PM, Selby PJ, Graham MA, et al. Pharmacokinetics of<br />

(Eq. 14)<br />

Parameter values used for the simulation were:<br />

pharmacokinetic parameters: kf = 5.73 h –1 , ke = 0.31<br />

h –1 , Vd = 80.96L; pharmacodynamic parameters for<br />

T helper cell suppression: CE50 = 9.2 µg • L –1 , Emax<br />

= 1; system parameters of T helper cell variation: kin<br />

= 383 cells • mm –3 • h –1 , kout = 0.335 h –1 , Rm =<br />

high dose methylprednisolone and use in hematological malignancies.<br />

Hematol Oncol 1993; 11 (2): 0.088 h –1 , Rb = 0.022 h –1 , tz = 13.2 (24-hour clock);<br />

89-96<br />

© 2005 Adis Data Information BV. All rights reserved. Clin Pharmacokinet 2005; 44 (1)


90 Czock et al.<br />

16. Al-Habet SM, Rogers HJ. Methylprednisolone pharmaco- 35. Rohatagi S, Barth J, Mollmann H, et al. Pharmacokinetics of<br />

kinetics after intravenous and oral administration. Br J Clin methylprednisolone and prednisolone after single and multiple<br />

Pharmacol 1989; 27 (3): 285-90 oral administration. J Clin Pharmacol 1997; 37 (10): 916-25<br />

17. Derendorf H, Mollmann H, Rohdewald P, et al. Kinetics of 36. Mollmann H, Hochhaus G, Rohatagi S, et al. Pharmacokinetic/<br />

methylprednisolone and its hemisuccinate ester. Clin pharmacodynamic evaluation of deflazacort in comparison to<br />

Pharmacol Ther 1985; 37 (5): 502-7 methylprednisolone and prednisolone. Pharm Res 1995; 12<br />

18. Rose JQ, Yurchak AM, Jusko WJ. Dose dependent pharmacokinetics<br />

(7): 1096-100<br />

of prednisone and prednisolone in man. J Pharma- 37. Ebling WF, Milsap RL, Szefler SJ, et al. 6 α-methylpred-<br />

cokinet Biopharm 1981; 9 (4): 389-417<br />

nisolone and 6 α-methylprednisone plasma protein binding in<br />

19. Tanner A, Bochner F, Caffin J, et al. Dose-dependent prednisolone<br />

humans and rabbits. J Pharm Sci 1986; 75 (8): 760-3<br />

kinetics. Clin Pharmacol Ther 1979; 25 (5 Pt 1): 571-8 38. Szefler SJ, Ebling WF, Georgitis JW, et al. Methylprednisolone<br />

20. Duggan DE, Yeh KC, Matalia N, et al. Bioavailability of oral versus prednisolone pharmacokinetics in relation to dose in<br />

dexamethasone. Clin Pharmacol Ther 1975; 18 (2): 205-9<br />

adults. Eur J Clin Pharmacol 1986; 30 (3): 323-9<br />

21. Rohatagi S, Hochhaus G, Mollmann H, et al. Pharmacokinetic 39. Szefler SJ, Rose JQ, Ellis EF, et al. The effect of troleandointeraction<br />

between endogenous cortisol and exogenous corti- mycin on methylprednisolone elimination. J Allergy Clin<br />

costeroids. Pharmazie 1995; 50 (9): 610-3<br />

Immunol 1980; 66 (6): 447-51<br />

22. Rosman PM, Benn R, Kay M, et al. Cortisol binding in uremic 40. Slayter KL, Ludwig EA, Lew KH, et al. Oral contraceptive<br />

plasma. I: absence of abnormal cortisol binding to cortisol effects on methylprednisolone pharmacokinetics and pharma-<br />

binding to corticosteroid-binding globulin. Nephron 1984; 37 codynamics. Clin Pharmacol Ther 1996; 59 (3): 312-21<br />

(3): 160-5 41. Cummings DM, Larijani GE, Conner DP, et al. Characteriza-<br />

23. Rosman PM, Benn R, Kay M, et al. Cortisol binding in uremic tion of dexamethasone binding in normal and uremic human<br />

plasma. II: decreased cortisol binding to albumin. Nephron serum. DICP 1990; 24 (3): 229-31<br />

1984; 37 (4): 229-31 42. Rohdewald P, Mollmann H, Barth J, et al. Pharmacokinetics of<br />

24. Wald JA, Law RM, Ludwig EA, et al. Evaluation of dose- dexamethasone and its phosphate ester. Biopharm Drug Disrelated<br />

pharmacokinetics and pharmacodynamics of predniso- pos 1987; 8 (3): 205-12<br />

lone in man. J Pharmacokinet Biopharm 1992; 20 (6): 567-89 43. Hochhaus G, Barth J, al-Fayoumi S, et al. Pharmacokinetics<br />

25. Dunn TE, Ludwig EA, Slaughter RL, et al. Pharmacokinetics and pharmacodynamics of dexamethasone sodium-m-suland<br />

pharmacodynamics of methylprednisolone in obesity. Clin fobenzoate (DS) after intravenous and intramuscular adminis-<br />

Pharmacol Ther 1991; 49 (5): 536-49<br />

tration: a comparison with dexamethasone phosphate (DP). J<br />

26. Ludwig EA, Slaughter RL, Savliwala M, et al. Steroid-specific Clin Pharmacol 2001; 41 (4): 425-34<br />

effects of ketoconazole on corticosteroid disposition: unaltered 44. Gupta SK, Ritchie JC, Ellinwood EH, et al. Modeling the<br />

prednisolone elimination. DICP 1989; 23 (11): 858-61<br />

pharmacokinetics and pharmacodynamics of dexamethasone<br />

27. Derendorf H, Rohdewald P, Mollmann H, et al. Pharmacokinetics<br />

in depressed patients. Eur J Clin Pharmacol 1992; 43 (1): 51-5<br />

of prednisolone after high doses of prednisolone 45. Lee KH, Shin JG, Chong WS, et al. Time course of the changes<br />

hemisuccinate. Biopharm Drug Dispos 1985; 6 (4): 423-32<br />

in prednisolone pharmacokinetics after co-administration or<br />

28. Meno-Tetang GM, Blum RA, Schwartz KE, et al. Effects of discontinuation of rifampin. Eur J Clin Pharmacol 1993; 45<br />

oral prasterone (dehydroepiandrosterone) on single-dose (3): 287-9<br />

pharmacokinetics of oral prednisone and cortisol suppression 46. Lebrun-Vignes B, Archer VC, Diquet B, et al. Effect of<br />

in normal women. J Clin Pharmacol 2001; 41 (11): 1195-205 itraconazole on the pharmacokinetics of prednisolone and<br />

29. Yamashita SK, Ludwig EA, Middleton Jr E, et al. Lack of methylprednisolone and cortisol secretion in healthy subjects.<br />

pharmacokinetic and pharmacodynamic interactions between Br J Clin Pharmacol 2001; 51 (5): 443-50<br />

ketoconazole and prednisolone. Clin Pharmacol Ther 1991; 49 47. Imani S, Jusko WJ, Steiner R. Diltiazem retards the metabolism<br />

(5): 558-70 of oral prednisone with effects on T-cell markers. Pediatr<br />

30. Booker BM, Magee MH, Blum RA, et al. Pharmacokinetic and Transplant 1999; 3 (2): 126-30<br />

pharmacodynamic interactions between diltiazem and methyl- 48. Kandrotas RJ, Slaughter RL, Brass C, et al. Ketoconazole<br />

prednisolone in healthy volunteers. Clin Pharmacol Ther 2002; effects on methylprednisolone disposition and their joint sup-<br />

72 (4): 370-82 pression of endogenous cortisol. Clin Pharmacol Ther 1987;<br />

31. Chow FS, Sharma A, Jusko WJ. Modeling interactions between 42 (4): 465-70<br />

adrenal suppression and T-helper lymphocyte trafficking 49. Glynn AM, Slaughter RL, Brass C, et al. Effects of ketoconaduring<br />

multiple dosing of methylprednisolone. J Pharma- zole on methylprednisolone pharmacokinetics and cortisol secokinet<br />

Biopharm 1999; 27 (6): 559-75 cretion. Clin Pharmacol Ther 1986; 39 (6): 654-9<br />

32. Barth J, Damoiseaux M, Mollmann H, et al. Pharmacokinetics 50. Varis T, Backman JT, Kivisto KT, et al. Diltiazem and<br />

and pharmacodynamics of prednisolone after intravenous and mibefradil increase the plasma concentrations and greatly enoral<br />

administration. Int J Clin Pharmacol Ther Toxicol 1992; hance the adrenal-suppressant effect of oral methylpred-<br />

30 (9): 317-24 nisolone. Clin Pharmacol Ther 2000; 67 (3): 215-21<br />

33. Bergrem H. The influence of uremia on pharmacokinetics and 51. Fost DA, Leung DY, Martin RJ, et al. Inhibition of methylprotein<br />

binding of prednisolone. Acta Med Scand 1983; 213 prednisolone elimination in the presence of clarithromycin<br />

(5): 333-7 therapy. J Allergy Clin Immunol 1999; 103 (6): 1031-5<br />

34. Boekenoogen SJ, Szefler SJ, Jusko WJ. Prednisolone disposi- 52. Varis T, Kivisto KT, Neuvonen PJ. Grapefruit juice can increase<br />

tion and protein binding in oral contraceptive users. J Clin the plasma concentrations of oral methylprednisolone. Eur J<br />

Endocrinol Metab 1983; 56 (4): 702-9 Clin Pharmacol 2000; 56 (6-7): 489-93<br />

© 2005 Adis Data Information BV. All rights reserved. Clin Pharmacokinet 2005; 44 (1)


Pharmacokinetics/Pharmacodynamics of Glucocorticoids 91<br />

53. Seidegard J, Simonsson M, Edsbacker S. Effect of an oral 70. Homma M, Oka K, Ikeshima K, et al. Different effects of<br />

contraceptive on the plasma levels of budesonide and prednis- traditional Chinese medicines containing similar herbal conolone<br />

and the influence on plasma cortisol. Clin Pharmacol stituents on prednisolone pharmacokinetics. J Pharm<br />

Ther 2000; 67 (4): 373-81 Pharmacol 1995; 47 (8): 687-92<br />

54. Tornatore KM, Morse GD, Jusko WJ, et al. Methylprednisolone 71. Luippold G, Schneider S, Marto M, et al. Pharmacokinetics of<br />

disposition in renal transplant recipients receiving triple-drug two oral prednisolone tablet formulations in healthy volunimmunosuppression.<br />

Transplantation 1989; 48 (6): 962-5 teers. Arzneimittel Forschung 2001; 51 (11): 911-5<br />

55. Milad MA, Ludwig EA, Lew KH, et al. The pharmacokinetics 72. Lewis GP, Jusko WJ, Graves L, et al. Prednisone side-effects<br />

and pharmacodynamics of methylprednisolone in chronic re- and serum-protein levels: a collaborative study. Lancet 1971;<br />

nal failure. Am J Ther 1994; 1 (1): 49-57 II (7728): 778-80<br />

56. Milsap RL, Plaisance KI, Jusko WJ. Prednisolone disposition in 73. Benet LZ, Hoener BA. Changes in plasma protein binding have<br />

obese men. Clin Pharmacol Ther 1984; 36 (6): 824-31 little clinical relevance. Clin Pharmacol Ther 2002; 71 (3):<br />

57. Wald JA, Jusko WJ. Corticosteroid pharmacodynamic model- 115-21<br />

ing: osteocalcin suppression by prednisolone. Pharm Res 74. Maser E, Volker B, Friebertshauser J. 11 β-hydroxysteroid<br />

1992; 9 (8): 1096-8 dehydrogenase type 1 from human liver: dimerization and<br />

58. Mager DE, Lin SX, Blum RA, et al. Dose equivalency evalua- enzyme cooperativity support its postulated role as glucocortition<br />

of major corticosteroids: pharmacokinetics and cell traf- coid reductase. Biochemistry 2002; 41 (7): 2459-65<br />

ficking and cortisol dynamics. J Clin Pharmacol 2003; 43 (11): 75. Diederich S, Hanke B, Burkhardt P, et al. Metabolism of<br />

1216-27 synthetic corticosteroids by 11 β-hydroxysteroid-dehydroge-<br />

59. Meibohm B, Derendorf H, Mollmann H, et al. Mechanism- nases in man. Steroids 1998; 63 (5-6): 271-7<br />

based PK/PD model for the lymphocytopenia induced by 76. Diederich S, Eigendorff E, Burkhardt P, et al. 11 β-hydroxysteendogenous<br />

and exogenous corticosteroids. Int J Clin roid dehydrogenase types 1 and 2: an important pharmacokine-<br />

Pharmacol Ther 1999; 37 (8): 367-76<br />

tic determinant for the activity of synthetic mineralo- and<br />

60. Oosterhuis B, ten Berge IJ, Schellekens PT, et al. Prednisolone glucocorticoids. J Clin Endocrinol Metab 2002; 87 (12): 5695-<br />

concentration-effect relations in humans and the influence of 701<br />

plasma hydrocortisone. J Pharmacol Exp Ther 1986; 239 (3):<br />

77. Esteban NV, Loughlin T, Yergey AL, et al. Daily cortisol<br />

919-26<br />

production rate in man determined by stable isotope dilution/<br />

61. Magee MH, Blum RA, Lates CD, et al. Pharmacokinetic/ mass spectrometry. J Clin Endocrinol Metab 1991; 72 (1): 39-<br />

pharmacodynamic model for prednisolone inhibition of whole 45<br />

blood lymphocyte proliferation. Br J Clin Pharmacol 2002; 53<br />

78. Lamberts SW, Bruining HA, de Jong FH. Corticosteroid therapy<br />

(5): 474-84<br />

in severe illness. N Engl J Med 1997; 337 (18): 1285-92<br />

62. Chakraborty A, Blum RA, Cutler DL, et al. Pharmacoimmuno-<br />

79. N’Gankam V, Uehlinger D, Dick B, et al. Increased cortisol<br />

dynamic interactions of interleukin-10 and prednisone in<br />

metabolites and reduced activity of 11 β-hydroxysteroid dehyhealthy<br />

volunteers. Clin Pharmacol Ther 1999; 65 (3): 304-18<br />

drogenase in patients on hemodialysis. Kidney Int 2002; 61<br />

63. Reiss WG, Slaughter RL, Ludwig EA, et al. Steroid dose<br />

(5): 1859-66<br />

sparing: pharmacodynamic responses to single versus divided<br />

80. Langhoff E, Olgaard K, Ladefoged J. The immunosuppressive<br />

doses of methylprednisolone in man. J Allergy Clin Immunol<br />

potency in vitro of physiological and synthetic steroids on<br />

1990; 85 (6): 1058-66<br />

lymphocyte cultures. Int J Immunopharmacol 1987; 9 (4): 469-<br />

64. Uhl A, Czock D, Boehm BO, et al. Pharmacokinetics and<br />

73<br />

pharmacodynamics of methylprednisolone after one bolus<br />

dose compared with two dose fractions. J Clin Pharm Ther<br />

81. Kraan GP, Dullaart RP, Pratt JJ, et al. Kinetics of intravenously<br />

2002; 27 (4): 281-7<br />

dosed cortisol in four men: consequences for calculation of the<br />

plasma cortisol production rate. J Steroid Biochem Mol Biol<br />

65. Derendorf H, Mollmann H, Krieg M, et al. Pharmacodynamics<br />

1997; 63 (1-3): 139-46<br />

of methylprednisolone phosphate after single intravenous administration<br />

to healthy volunteers. Pharm Res 1991; 8 (2): 82. Boudinot FD, Jusko WJ. Plasma protein binding interaction of<br />

263-8<br />

prednisone and prednisolone. J Steroid Biochem 1984; 21 (3):<br />

66. Braat MC, Oosterhuis B, Koopmans RP, et al. Kinetic-dynamic<br />

337-9<br />

modeling of lymphocytopenia induced by the combined action 83. Frey BM, Frey FJ. Clinical pharmacokinetics of prednisone and<br />

of dexamethasone and hydrocortisone in humans, after inhala- prednisolone. Clin Pharmacokinet 1990; 19 (2): 126-46<br />

tion and intravenous administration of dexamethasone. J 84. Magee MH, Blum RA, Lates CD, et al. Prednisolone pharmaco-<br />

Pharmacol Exp Ther 1992; 262 (2): 509-15<br />

kinetics and pharmacodynamics in relation to sex and race. J<br />

67. Oosterhuis B, ten Berge RJ, Sauerwein HP, et al. Pharmacokin- Clin Pharmacol 2001; 41 (11): 1180-94<br />

etic-pharmacodynamic modeling of prednisolone-induced 85. Bartoszek M, Brenner AM, Szefler SJ. Prednisolone and<br />

lymphocytopenia in man. J Pharmacol Exp Ther 1984; 229 (2): methylprednisolone kinetics in children receiving anticonvul-<br />

539-46<br />

sant therapy. Clin Pharmacol Ther 1987; 42 (4): 424-32<br />

68. Lew KH, Ludwig EA, Milad MA, et al. Gender-based effects 86. Wassner SJ, Malekzadeh MH, Pennisi AJ, et al. Allograft<br />

on methylprednisolone pharmacokinetics and pharmaco- survival in patients receiving anticonvulsant medications. Clin<br />

dynamics. Clin Pharmacol Ther 1993; 54 (4): 402-14<br />

Nephrol 1977; 8 (1): 293-7<br />

69. Tornatore KM, Logue G, Venuto RC, et al. Pharmacokinetics 87. Buffington GA, Dominguez JH, Piering WF, et al. Interaction<br />

of methylprednisolone in elderly and young healthy males. J of rifampin and glucocorticoids: adverse effect on renal al-<br />

Am Geriatr Soc 1994; 42 (10): 1118-22 lograft function. JAMA 1976; 236 (17): 1958-60<br />

© 2005 Adis Data Information BV. All rights reserved. Clin Pharmacokinet 2005; 44 (1)


92 Czock et al.<br />

88. Hollander AA, van Rooij J, Lentjes GW, et al. The effect of 107. Villikka K, Kivisto KT, Neuvonen PJ. The effect of dexamethgrapefruit<br />

juice on cyclosporine and prednisone metabolism in asone on the pharmacokinetics of triazolam. Pharmacol Toxtransplant<br />

patients. Clin Pharmacol Ther 1995; 57 (3): 318-24 icol 1998; 83 (3): 135-8<br />

89. Jeng S, Chanchairujira T, Jusko W, et al. Prednisone metabol- 108. Villikka K, Varis T, Backman JT, et al. Effect of methylism<br />

in recipients of kidney or liver transplants and in lung prednisolone on CYP3A4-mediated drug metabolism in vivo.<br />

recipients receiving ketoconazole. Transplantation 2003; 75 Eur J Clin Pharmacol 2001; 57 (6-7): 457-60<br />

(6): 792-5 109. Yates CR, Vysokanov A, Mukherjee A, et al. Time-variant<br />

90. Ost L. Impairment of prednisolone metabolism by cyclosporine increase in methylprednisolone clearance in patients with<br />

treatment in renal graft recipients. Transplantation 1987; 44 acute respiratory distress syndrome: a population pharmaco-<br />

(4): 533-5 kinetic study. J Clin Pharmacol 2001; 41 (4): 415-24<br />

91. Jusko WJ, Ferron GM, Mis SM, et al. Pharmacokinetics of 110. Shimada T, Terada A, Yokogawa K, et al. Lowered blood<br />

prednisolone during administration of sirolimus in patients concentration of tacrolimus and its recovery with changes in<br />

with renal transplants. J Clin Pharmacol 1996; 36 (12): 1100-6 expression of CYP3A and P-glycoprotein after high-dose ster-<br />

92. Ludwig EA, Kong AN, Camara DS, et al. Pharmacokinetics of oid therapy. Transplantation 2002; 74 (10): 1419-24<br />

methylprednisolone hemisuccinate and methylprednisolone in 111. Lackner TE. Interaction of dexamethasone with phenytoin.<br />

chronic liver disease. J Clin Pharmacol 1993; 33 (9): 805-10 Pharmacotherapy 1991; 11 (4): 344-7<br />

93. Periti P, Mazzei T, Mini E, et al. Pharmacokinetic drug interac- 112. Cattaneo D, Perico N, Gaspari F, et al. Glucocorticoids interfere<br />

tions of macrolides. Clin Pharmacokinet 1992; 23 (2): 106-31 with mycophenolate mofetil bioavailability in kidney trans-<br />

94. LaForce CF, Szefler SJ, Miller MF, et al. Inhibition of methyl- plantation. Kidney Int 2002; 62 (3): 1060-7<br />

prednisolone elimination in the presence of erythromycin ther-<br />

113. Schuetz EG, Hazelton GA, Hall J, et al. Induction of digitoxapy.<br />

J Allergy Clin Immunol 1983; 72 (1): 34-9<br />

igenin monodigitoxoside UDP-glucuronosyltransferase ac-<br />

95. Szefler SJ, Brenner M, Jusko WJ, et al. Dose- and time-related tivity by glucocorticoids and other inducers of cytochrome P-<br />

effect of troleandomycin on methylprednisolone elimination. 450p in primary monolayer cultures of adult rat hepatocytes<br />

Clin Pharmacol Ther 1982; 32 (2): 166-71 and in human liver. J Biol Chem 1986; 261 (18): 8270-5<br />

96. Kamada AK, Hill MR, Ikle DN, et al. Efficacy and safety of<br />

114. Frey FJ, Schnetzer A, Horber FF, et al. Evidence that cyclolow-dose<br />

troleandomycin therapy in children with severe, stersporine<br />

does not affect the metabolism of prednisolone after<br />

oid-requiring asthma. J Allergy Clin Immunol 1993; 91 (4):<br />

renal transplantation. Transplantation 1987; 43 (4): 494-8<br />

873-82<br />

115. Rocci Jr ML, Tietze KJ, Lee J, et al. The effect of cyclosporine<br />

97. Zeiger RS, Schatz M, Sperling W, et al. Efficacy of troleandoon<br />

the pharmacokinetics of prednisolone in renal transplant<br />

mycin in outpatients with severe, corticosteroid-dependent<br />

patients. Transplantation 1988; 45 (3): 656-60<br />

asthma. J Allergy Clin Immunol 1980; 66 (6): 438-46<br />

116. Ptachcinski RJ, Venkataramanan R, Burckart GJ, et al. Cyclo-<br />

98. Sirgo MA, Rocci Jr ML, Ferguson RK, et al. Effects of<br />

sporine: high-dose steroid interaction in renal transplant recipicimetidine<br />

and ranitidine on the conversion of prednisone to<br />

ents: assessment by HPLC. Transplant Proc 1987; 19 (1 Pt 2):<br />

prednisolone. Clin Pharmacol Ther 1985; 37 (5): 534-8<br />

1728-9<br />

99. Green AW, Ebling WF, Gardner MJ, et al. Cimetidine-methylprednisolone-theophylline<br />

metabolic interaction. Am J Med<br />

117. Klintmalm G, Sawe J, Ringden O, et al. Cyclosporine plasma<br />

1984; 77 (6): 1115-8<br />

levels in renal transplant patients: association with renal toxicity<br />

and allograft rejection. Transplantation 1985; 39 (2): 132-7<br />

100. Yates CR, Chang C, Kearbey JD, et al. Structural determinants<br />

of P-glycoprotein-mediated transport of glucocorticoids. 118. Arnold JC, O’Grady JG, Tredger JM, et al. Effects of low-dose<br />

Pharm Res 2003; 20 (11): 1794-803<br />

prednisolone on cyclosporine pharmacokinetics in liver trans-<br />

101. Karssen AM, Meijer OC, van der Sandt IC, et al. The role of the<br />

plant recipients: radioimmunoassay with specific and nonefflux<br />

transporter P-glycoprotein in brain penetration of predspecific<br />

monoclonal antibodies. Eur J Clin Pharmacol 1990; 39<br />

nisolone. J Endocrinol 2002; 175 (1): 251-60<br />

(3): 257-60<br />

102. Koszdin KL, Shen DD, Bernards CM. Spinal cord bioavailabiltrough<br />

concentrations are not affected by bolus methyl-<br />

119. Backman L, Kreis H, Morales JM, et al. Sirolimus steady-state<br />

ity of methylprednisolone after intravenous and intrathecal<br />

administration: the role of P-glycoprotein. Anesthesiology prednisolone therapy in renal allograft recipients. Br J Clin<br />

2000; 92 (1): 156-63<br />

Pharmacol 2002; 54 (1): 65-8<br />

103. Kovarik JM, Purba HS, Pongowski M, et al. Pharmacokinetics 120. Chamberlain CE, Kirk AD, Hale D, et al. Prednisone decreases<br />

of dexamethasone and valspodar, a P-glycoprotein (MDR1) sensitivity to sirolimus in kidney transplant recipients [ab-<br />

modulator: implications for coadministration. Pharmacother- stract]. Am J Transplant 2003; 3 Suppl. 5: 354<br />

apy 1998; 18 (6): 1230-6<br />

121. Frey FJ, Gambertoglio JG, Frey BM, et al. Nonlinear plasma<br />

104. Rose JQ, Yurchak AM, Meikle AW, et al. Effect of smoking on protein binding and haemodialysis clearance of prednisolone.<br />

prednisone, prednisolone, and dexamethasone pharmaco- Eur J Clin Pharmacol 1982; 23 (1): 65-74<br />

kinetics. J Pharmacokinet Biopharm 1981; 9 (1): 1-14 122. Sherlock JE, Letteri JM. Effect of hemodialysis on methyl-<br />

105. Pascussi JM, Gerbal-Chaloin S, Drocourt L, et al. The expres- prednisolone plasma levels. Nephron 1977; 18 (4): 208-11<br />

sion of CYP2B6, CYP2C9 and CYP3A4 genes: a tangle of 123. Bergrem H. Pharmacokinetics and protein binding of prednisonetworks<br />

of nuclear and steroid receptors. Biochim Biophys lone in patients with nephrotic syndrome and patients undergo-<br />

Acta 2003; 1619 (3): 243-53<br />

ing hemodialysis. Kidney Int 1983; 23 (6): 876-81<br />

106. McCune JS, Hawke RL, LeCluyse EL, et al. In vivo and in vitro 124. Schwab M, Klotz U. Pharmacokinetic considerations in the<br />

induction of human cytochrome P4503a4 by dexamethasone. treatment of inflammatory bowel disease. Clin Pharmacokinet<br />

Clin Pharmacol Ther 2000; 68 (4): 356-66 2001; 40 (10): 723-51<br />

© 2005 Adis Data Information BV. All rights reserved. Clin Pharmacokinet 2005; 44 (1)


Pharmacokinetics/Pharmacodynamics of Glucocorticoids 93<br />

125. Milsap RL, George DE, Szefler SJ, et al. Effect of inflammatory 143. Bodwell JE, Webster JC, Jewell CM, et al. Glucocorticoid<br />

bowel disease on absorption and disposition of prednisolone. receptor phosphorylation: overview, function and cell cycle-<br />

Dig Dis Sci 1983; 28 (2): 161-8 dependence. J Steroid Biochem Mol Biol 1998; 65 (1-6): 91-9<br />

126. Dove AM, Szefler SJ, Hill MR, et al. Altered prednisolone 144. Irusen E, Matthews JG, Takahashi A, et al. P38 mitogenpharmacokinetics<br />

in patients with cystic fibrosis. J Pediatr activated protein kinase-induced glucocorticoid receptor phos-<br />

1992; 120 (5): 789-94<br />

phorylation reduces its activity: role in steroid-insensitive asth-<br />

127. Rose JQ, Nickelsen JA, Middleton Jr E, et al. Prednisolone ma. J Allergy Clin Immunol 2002; 109 (4): 649-57<br />

disposition in steroid-dependent asthmatics. J Allergy Clin 145. DeFranco DB. Navigating steroid hormone receptors through<br />

Immunol 1980; 66 (5): 366-73<br />

the nuclear compartment. Mol Endocrinol 2002; 16 (7): 1449-<br />

128. Petersen KB, Jusko WJ, Rasmussen M, et al. Population<br />

55<br />

pharmacokinetics of prednisolone in children with acute lymdexamethasone<br />

and RU486 on glucocorticoid receptor-DNA<br />

146. Pandit S, Geissler W, Harris G, et al. Allosteric effects of<br />

phoblastic leukemia. Cancer Chemother Pharmacol 2003; 51<br />

(6): 465-73<br />

interactions. J Biol Chem 2002; 277 (2): 1538-43<br />

147. Denk A, Goebeler M, Schmid S, et al. Activation of NF-κB via<br />

129. Segal JL, Maltby BF, Langdorf MI, et al. Methylprednisolone<br />

the IκB kinase complex is both essential and sufficient for<br />

disposition kinetics in patients with acute spinal cord injury.<br />

proinflammatory gene expression in primary endothelial cells.<br />

Pharmacotherapy 1998; 18 (1): 16-22<br />

J Biol Chem 2001; 276 (30): 28451-8<br />

130. Newton R. Molecular mechanisms of glucocorticoid action:<br />

148. Vermeulen L, De Wilde G, Notebaert S, et al. Regulation of the<br />

what is important? Thorax 2000; 55 (7): 603-13<br />

transcriptional activity of the nuclear factor-κB p65 subunit.<br />

131. Buttgereit F, Wehling M, Burmester GR. A new hypothesis of Biochem Pharmacol 2002; 64 (5-6): 963-70<br />

modular glucocorticoid actions: steroid treatment of rheumatic 149. Lawrence T, Willoughby DA, Gilroy DW. Anti-inflammatory<br />

diseases revisited. Arthritis Rheum 1998; 41 (5): 761-7<br />

lipid mediators and insights into the resolution of inflamma-<br />

132. Colbert RA, Young DA. Glucocorticoid-induced messenger tion. Nat Rev Immunol 2002; 2 (10): 787-95<br />

ribonucleic acids in rat thymic lymphocytes: rapid primary 150. Kagoshima M, Ito K, Cosio B, et al. Glucocorticoid suppression<br />

effects specific for glucocorticoids. Endocrinology 1986; 119 of nuclear factor-κB: a role for histone modifications. Bio-<br />

(6): 2598-605 chem Soc Trans 2003; 31 Pt 1: 60-5<br />

133. Jin JY, Almon RR, DuBois DC, et al. Modeling of corticoster- 151. Ito K, Barnes PJ, Adcock IM. Glucocorticoid receptor recruitoid<br />

pharmacogenomics in rat liver using gene microarrays. J ment of histone deacetylase 2 inhibits interleukin-1 β-induced<br />

Pharmacol Exp Ther 2003; 307 (1): 93-109<br />

histone H4 acetylation on lysines 8 and 12. Mol Cell Biol<br />

134. Almon RR, DuBois DC, Brandenburg EH, et al. Pharmaco- 2000; 20 (18): 6891-903<br />

dynamics and pharmacogenomics of diverse receptor-meditranscriptional<br />

152. Scheinman RI, Cogswell PC, Lofquist AK, et al. Role of<br />

ated effects of methylprednisolone in rats using microarray<br />

activation of IκBα in mediation of immunosup-<br />

analysis. J Pharmacokinet Pharmacodyn 2002; 29 (2): 103-29 pression by glucocorticoids. Science 1995; 270 (5234): 283-6<br />

135. Galon J, Franchimont D, Hiroi N, et al. Gene profiling reveals 153. Poon M, Liu B, Taubman MB. Identification of a novel dexaunknown<br />

enhancing and suppressive actions of glucocorticyte<br />

methasone-sensitive RNA-destabilizing region on rat mono-<br />

coids on immune cells. FASEB J 2002; 16 (1): 61-71<br />

chemoattractant protein 1 mRNA. Mol Cell Biol 1999; 19<br />

136. Croxtall JD, Choudhury Q, Flower RJ. Glucocorticoids act<br />

(10): 6471-8<br />

within minutes to inhibit recruitment of signalling factors to<br />

154. Newton R, Seybold J, Kuitert LM, et al. Repression of cyclo-<br />

activated EGF receptors through a receptor-dependent, tranoccurs<br />

by transcriptional and post-transcriptional mechanisms<br />

oxygenase-2 and prostaglandin E2 release by dexamethasone<br />

scription-independent mechanism. Br J Pharmacol 2000; 130<br />

(2): 289-98<br />

involving loss of polyadenylated mRNA. J Biol Chem 1998;<br />

273 (48): 32312-21<br />

137. Tanioka T, Nakatani Y, Semmyo N, et al. Molecular identifica-<br />

155. Hammes SR. The further redefining of steroid-mediated signaltion<br />

of cytosolic prostaglandin E2 synthase that is functionally<br />

ing. Proc Natl Acad Sci USA 2003; 100 (5): 2168-70<br />

coupled with cyclooxygenase-1 in immediate prostaglandin E2<br />

biosynthesis. J Biol Chem 2000; 275 (42): 32775-82<br />

156. Zhu Y, Rice CD, Pang Y, et al. Cloning, expression, and<br />

characterization of a membrane progestin receptor and evi-<br />

138. Ylikomi T, Wurtz JM, Syvala H, et al. Reappraisal of the role of<br />

dence it is an intermediary in meiotic maturation of fish<br />

heat shock proteins as regulators of steroid receptor activity.<br />

oocytes. Proc Natl Acad Sci USA 2003; 100 (5): 2231-6<br />

Crit Rev Biochem Mol Biol 1998; 33 (6): 437-66<br />

157. Buttgereit F, Scheffold A. Rapid glucocorticoid effects on<br />

139. Lieberman BA, Nordeen SK. DNA intersegment transfer, how immune cells. Steroids 2002; 67 (6): 529-34<br />

steroid receptors search for a target site. J Biol Chem 1997; 158. Hinz B, Hirschelmann R. Rapid non-genomic feedback effects<br />

272 (2): 1061-8 of glucocorticoids on CRF-induced ACTH secretion in rats.<br />

140. Evans RM. The steroid and thyroid hormone receptor superfam- Pharm Res 2000; 17 (10): 1273-7<br />

ily. Science 1988; 240 (4854): 889-95<br />

159. Falkenstein E, Tillmann HC, Christ M, et al. Multiple actions of<br />

141. Rogatsky I, Luecke HF, Leitman DC, et al. Alternate surfaces steroid hormones: a focus on rapid, nongenomic effects.<br />

of transcriptional coregulator GRIP1 function in different Pharmacol Rev 2000; 52 (4): 513-56<br />

glucocorticoid receptor activation and repression contexts. 160. Croxtall JD, van Hal PT, Choudhury Q, et al. Different gluco-<br />

Proc Natl Acad Sci USA 2002; 99 (26): 16701-6<br />

corticoids vary in their genomic and non-genomic mechanism<br />

142. Rosenfeld MG, Glass CK. Coregulator codes of transcriptional of action in A549 cells. Br J Pharmacol 2002; 135 (2): 511-9<br />

regulation by nuclear receptors. J Biol Chem 2001; 276 (40): 161. Buttgereit F, Brand MD, Burmester GR. Equivalent doses and<br />

36865-8 relative drug potencies for non-genomic glucocorticoid ef-<br />

© 2005 Adis Data Information BV. All rights reserved. Clin Pharmacokinet 2005; 44 (1)


94 Czock et al.<br />

fects: a novel glucocorticoid hierarchy. Biochem Pharmacol 179. Filaci G, Suciu-Foca N. CD8+ T suppressor cells are back to the<br />

1999; 58 (2): 363-8 game: are they players in autoimmunity? Autoimmun Rev<br />

162. Croxtall JD, Gilroy DW, Solito E, et al. Attenuation of 2002; 1 (5): 279-83<br />

glucocorticoid functions in an ANX-A1 -/- cell line. Biochem J 180. Santana MA, Rosenstein Y. What it takes to become an effector<br />

2003; 371 Pt 3: 927-35 T cell: the process, the cells involved, and the mechanisms. J<br />

163. Whiting KP, Restall CJ, Brain PF. Steroid hormone-induced Cell Physiol 2003; 195 (3): 392-401<br />

effects on membrane fluidity and their potential roles in non- 181. Lechler RI, Garden OA, Turka LA. The complementary roles of<br />

genomic mechanisms. Life Sci 2000; 67 (7): 743-57<br />

deletion and regulation in transplantation tolerance. Nat Rev<br />

164. Lamche HR, Silberstein PT, Knabe AC, et al. Steroids decrease Immunol 2003; 3 (2): 147-58<br />

granulocyte membrane fluidity, while phorbol ester increases 182. Chen W, Wahl SM. TGF-β: the missing link in CD4(+)CD25(+)<br />

membrane fluidity: studies using electron paramagnetic regulatory T cell-mediated immunosuppression. Cytokine<br />

resonance. Inflammation 1990; 14 (1): 61-70 Growth Factor Rev 2003; 14 (2): 85-9<br />

165. Hinz B, Hirschelmann R. Dexamethasone megadoses stabilize 183. Kalthoff FS, Chung J, Musser P, et al. Pimecrolimus does not<br />

rat liver lysosomal membranes by non-genomic and genomic affect the differentiation, maturation and function of human<br />

effects. Pharm Res 2000; 17 (12): 1489-93<br />

monocyte-derived dendritic cells, in contrast to corticosteroids.<br />

166. Almawi WY, Beyhum HN, Rahme AA, et al. Regulation of<br />

Clin Exp Immunol 2003; 133 (3): 350-9<br />

cytokine and cytokine receptor expression by glucocorticoids. 184. Rea D, van Kooten C, van Meijgaarden KE, et al. Glucocorti-<br />

J Leukoc Biol 1996; 60 (5): 563-72<br />

coids transform CD40-triggering of dendritic cells into an<br />

167. Khan WA, Blobe GC, Hannun YA. Arachidonic acid and free alternative activation pathway resulting in antigen-presenting<br />

fatty acids as second messengers and the role of protein kinase cells that secrete IL-10. Blood 2000; 95 (10): 3162-7<br />

C. Cell Signal 1995; 7 (3): 171-84 185. Dong X, Bachman LA, Kumar R, et al. Generation of antigenspecific,<br />

168. Newton R, Kuitert LM, Slater DM, et al. Cytokine induction of<br />

interleukin-10-producing T-cells using dendritic cell<br />

cytosolic phospholipase a2 and cyclooxygenase-2 mRNA is stimulation and steroid hormone conditioning. Transpl<br />

suppressed by glucocorticoids in human epithelial cells. Life Immunol 2003; 11 (3-4): 323-33<br />

Sci 1997; 60 (1): 67-78<br />

186. Barrat FJ, Cua DJ, Boonstra A, et al. In vitro generation of<br />

169. Matsumura M, Kakishita H, Suzuki M, et al. Dexamethasone interleukin 10-producing regulatory CD4(+) T cells is induced<br />

suppresses iNOS gene expression by inhibiting NF-κB in by immunosuppressive drugs and inhibited by T helper type 1<br />

vascular smooth muscle cells. Life Sci 2001; 69 (9): 1067-77 (Th1)- and Th2-inducing cytokines. J Exp Med 2002; 195 (5):<br />

170. Shinoda J, McLaughlin KE, Bell HS, et al. Molecular mechan- 603-16<br />

isms underlying dexamethasone inhibition of iNOS expression 187. Terada Y, Okado T, Inoshita S, et al. Glucocorticoids stimulate<br />

and activity in C6 glioma cells. Glia 2003; 42 (1): 68-76<br />

p21(CIP1) in mesangial cells and in anti-GBM glomerulone-<br />

171. Linnane SJ, Thin AG, Keatings VM, et al. Glucocorticoid phritis. Kidney Int 2001; 59 (5): 1706-16<br />

treatment reduces exhaled nitric oxide in cystic fibrosis pacoid<br />

188. Goke MN, Schneider M, Beil W, et al. Differential glucocorti-<br />

tients. Eur Respir J 2001; 17 (6): 1267-70<br />

effects on repair mechanisms and NF-κB activity in the<br />

172. Moodley YP, Chetty R, Lalloo UG. Nitric oxide levels in intestinal epithelium. Regul Pept 2002; 105 (3): 203-14<br />

exhaled air and inducible nitric oxide synthase immunolocalapoptosis<br />

189. Amsterdam A, Tajima K, Sasson R. Cell-specific regulation of<br />

ization in pulmonary sarcoidosis. Eur Respir J 1999; 14 (4):<br />

by glucocorticoids: implication to their anti-inflam-<br />

822-7<br />

matory action. Biochem Pharmacol 2002; 64 (5-6): 843-50<br />

173. Pitzalis C, Pipitone N, Perretti M. Regulation of leukocyteprednisolone<br />

190. Leussink VI, Jung S, Merschdorf U, et al. High-dose methylendothelial<br />

interactions by glucocorticoids. Ann N Y Acad Sci<br />

therapy in multiple sclerosis induces apoptosis in<br />

2002; 966: 108-18<br />

peripheral blood leukocytes. Arch Neurol 2001; 58 (1): 91-7<br />

174. Atsuta J, Plitt J, Bochner BS, et al. Inhibition of VCAM-1 191. Migita K, Eguchi K, Kawabe Y, et al. Apoptosis induction in<br />

expression in human bronchial epithelial cells by glucocortitherapy.<br />

human peripheral blood T lymphocytes by high-dose steroid<br />

coids. Am J Respir Cell Mol Biol 1999; 20 (4): 643-50<br />

Transplantation 1997; 63 (4): 583-7<br />

175. Cronstein BN, Kimmel SC, Levin RI, et al. A mechanism for 192. Schmidt J, Gold R, Schonrock L, et al. T-cell apoptosis in situ<br />

the antiinflammatory effects of corticosteroids: the glucocortimethylprednisolone<br />

in experimental autoimmune encephalomyelitis following<br />

coid receptor regulates leukocyte adhesion to endothelial cells<br />

pulse therapy. Brain 2000; 123 Pt 7: 1431-<br />

and expression of endothelial-leukocyte adhesion molecule 1 41<br />

and intercellular adhesion molecule 1. Proc Natl Acad Sci 193. Zhang X, Moilanen E, Adcock IM, et al. Divergent effect of<br />

USA 1992; 89 (21): 9991-5<br />

mometasone on human eosinophil and neutrophil apoptosis.<br />

176. Goulding NJ, Ogbourn S, Pipitone N, et al. The inhibitory effect Life Sci 2002; 71 (13): 1523-34<br />

of dexamethasone on lymphocyte adhesion molecule expres- 194. Meagher LC, Cousin JM, Seckl JR, et al. Opposing effects of<br />

sion and intercellular aggregation is not mediated by lipocortin glucocorticoids on the rate of apoptosis in neutrophilic and<br />

1. Clin Exp Immunol 1999; 118 (3): 376-83 eosinophilic granulocytes. J Immunol 1996; 156 (11): 4422-8<br />

177. Bluestone JA, Abbas AK. Natural versus adaptive regulatory T 195. Baichwal VR, Baeuerle PA. Activate NF-κB or die? Curr Biol<br />

cells. Nat Rev Immunol 2003; 3 (3): 253-7 1997; 7 (2): R94-6<br />

178. Levings MK, Sangregorio R, Sartirana C, et al. Human 196. Stankova J, Turcotte S, Harris J, et al. Modulation of leuko-<br />

CD25+CD4+ T suppressor cell clones produce transforming triene B4 receptor-1 expression by dexamethasone: potential<br />

growth factor β, but not interleukin 10, and are distinct from mechanism for enhanced neutrophil survival. J Immunol 2002;<br />

type 1 T regulatory cells. J Exp Med 2002; 196 (10): 1335-46 168 (7): 3570-6<br />

© 2005 Adis Data Information BV. All rights reserved. Clin Pharmacokinet 2005; 44 (1)


Pharmacokinetics/Pharmacodynamics of Glucocorticoids 95<br />

197. Webster JC, Huber RM, Hanson RL, et al. Dexamethasone and 215. Schlaghecke R, Kornely E, Santen RT, et al. The effect of longtumor<br />

necrosis factor-α act together to induce the cellular term glucocorticoid therapy on pituitary-adrenal responses to<br />

inhibitor of apoptosis-2 gene and prevent apoptosis in a variety exogenous corticotropin-releasing hormone. N Engl J Med<br />

of cell types. Endocrinology 2002; 143 (10): 3866-74 1992; 326 (4): 226-30<br />

198. Kuroda M, Sasamura H, Shimizu-Hirota R, et al. Glucocortipatients<br />

216. Agusti C, Rano A, Filella X, et al. Pulmonary infiltrates in<br />

coid regulation of proteoglycan synthesis in mesangial cells.<br />

receiving long-term glucocorticoid treatment: etiolo-<br />

Kidney Int 2002; 62 (3): 780-9<br />

gy, prognostic factors, and associated inflammatory response.<br />

199. Kasinath BS, Singh AK, Kanwar YS, et al. Dexamethasone Chest 2003; 123 (2): 488-98<br />

increases heparan sulfate proteoglycan core protein content of 217. Porges AJ, Beattie SL, Ritchlin C, et al. Patients with systemic<br />

glomerular epithelial cells. J Lab Clin Med 1990; 115 (2): 196- lupus erythematosus at risk for pneumocystis carinii pneumo-<br />

202<br />

nia. J Rheumatol 1992; 19 (8): 1191-4<br />

200. Russell RE, Culpitt SV, DeMatos C, et al. Release and activity 218. Canalis E, Delany AM. Mechanisms of glucocorticoid action in<br />

of matrix metalloproteinase-9 and tissue inhibitor of metal- bone. Ann N Y Acad Sci 2002; 966: 73-81<br />

loproteinase-1 by alveolar macrophages from patients with 219. Engelbrecht Y, de Wet H, Horsch K, et al. Glucocorticoids<br />

chronic obstructive pulmonary disease. Am J Respir Cell Mol induce rapid up-regulation of mitogen-activated protein kinase<br />

Biol 2002; 26 (5): 602-9<br />

phosphatase-1 and dephosphorylation of extracellular signal-<br />

201. Perretti M, Ahluwalia A. The microcirculation and inflammaosteoblast<br />

regulated kinase and impair proliferation in human and mouse<br />

tion: site of action for glucocorticoids. Microcirculation 2000;<br />

cell lines. Endocrinology 2003; 144 (2): 412-22<br />

7 (3): 147-61<br />

220. Van Staa TP, Leufkens HG, Abenhaim L, et al. Use of oral<br />

202. Perretti M, Chiang N, La M, et al. Endogenous lipid- and corticosteroids and risk of fractures. J Bone Miner Res 2000;<br />

peptide-derived anti-inflammatory pathways generated with 15 (6): 993-1000<br />

glucocorticoid and aspirin treatment activate the lipoxin A4 221. van Staa TP, Leufkens HG, Abenhaim L, et al. Oral corticoster-<br />

receptor. Nat Med 2002; 8 (11): 1296-302<br />

oids and fracture risk: relationship to daily and cumulative<br />

203. Stelmach I, Jerzynska J, Kuna P. A randomized, double-blind doses. Rheumatology (Oxf) 2000; 39 (12): 1383-9<br />

trial of the effect of glucocorticoid, antileukotriene and β- 222. Garcia Rodriguez LA, Hernandez-Diaz S. The risk of upper<br />

agonist treatment on IL-10 serum levels in children with gastrointestinal complications associated with nonsteroidal<br />

asthma. Clin Exp Allergy 2002; 32 (2): 264-9<br />

anti-inflammatory drugs, glucocorticoids, acetaminophen, and<br />

204. Asadullah K, Sterry W, Volk HD. Interleukin-10 therapy: recombinations<br />

of these agents. Arthritis Res 2001; 3 (2): 98-101<br />

view of a new approach. Pharmacol Rev 2003; 55 (2): 241-69<br />

223. Guslandi M, Tittobello A. Steroid ulcers: a myth revisited. BMJ<br />

1992; 304 (6828): 655-6<br />

205. Segerer S, Cui Y, Eitner F, et al. Expression of chemokines and<br />

chemokine receptors during human renal transplant rejection.<br />

224. Jobling AI, Augusteyn RC. What causes steroid cataracts? A<br />

Am J Kidney Dis 2001; 37 (3): 518-31<br />

review of steroid-induced posterior subcapsular cataracts. Clin<br />

Exp Optom 2002; 85 (2): 61-75<br />

206. Game DS, Lechler RI. Pathways of allorecognition: implica-<br />

225. Vreden SG, Hermus AR, van Liessum PA, et al. Aseptic bone<br />

tions for transplantation tolerance. Transpl Immunol 2002; 10<br />

necrosis in patients on glucocorticoid replacement therapy.<br />

(2-3): 101-8<br />

Neth J Med 1991; 39 (3-4): 153-7<br />

207. Pan J, Ju D, Wang Q, et al. Dexamethasone inhibits the antigen<br />

226. Ahmed SF, Tucker P, Mushtaq T, et al. Short-term effects on<br />

presentation of dendritic cells in MHC class II pathway.<br />

linear growth and bone turnover in children randomized to<br />

Immunol Lett 2001; 76 (3): 153-61<br />

receive prednisolone or dexamethasone. Clin Endocrinol (Oxf)<br />

208. Almawi WY, Lipman ML, Stevens AC, et al. Abrogation of 2002; 57 (2): 185-91<br />

glucocorticoid-mediated inhibition of t cell proliferation by the<br />

227. Ventura MT, Calogiuri GF, Matino MG, et al. Alternative<br />

synergistic action of IL-1, IL-6, and IFN-γ. J Immunol 1991;<br />

glucocorticoids for use in cases of adverse reaction to systemic<br />

146 (10): 3523-7<br />

glucocorticoids: a study on 10 patients. Br J Dermatol 2003;<br />

209. Park SY, Kim HW, Moon KC, et al. mRNA expression of 148 (1): 139-41<br />

intercellular adhesion molecule-1 and vascular cell adhesion 228. Twycross R. The risks and benefits of corticosteroids in admolecule-1<br />

in acute renal allograft rejection. Transplantation vanced cancer. Drug Saf 1994; 11 (3): 163-78<br />

2000; 69 (12): 2554-60<br />

229. Wilson AM, Lipworth BJ. Short-term dose-response relation-<br />

210. Frauman AG. An overview of the adverse reactions to adrenal ships for the relative systemic effects of oral prednisolone and<br />

corticosteroids. Adverse Drug React Toxicol Rev 1996; 15 (4): inhaled fluticasone in asthmatic adults. Br J Clin Pharmacol<br />

203-6 1999; 48 (4): 579-85<br />

211. Schacke H, Docke WD, Asadullah K. Mechanisms involved in 230. Vermeer H, Hendriks-Stegeman BI, van der Burg B, et al.<br />

the side effects of glucocorticoids. Pharmacol Ther 2002; 96 Glucocorticoid-induced increase in lymphocytic FKBP51<br />

(1): 23-43 messenger ribonucleic acid expression: a potential marker for<br />

212. Whitworth JA, Gordon D, Andrews J, et al. The hypertensive glucocorticoid sensitivity, potency, and bioavailability. J Clin<br />

effect of synthetic glucocorticoids in man: role of sodium and Endocrinol Metab 2003; 88 (1): 277-84<br />

volume. J Hypertens 1989; 7 (7): 537-49<br />

231. Jaffuel D, Roumestan C, Balaguer P, et al. Correlation between<br />

213. Brem AS. Insights into glucocorticoid-associated hypertension. different gene expression assays designed to measure trans-<br />

Am J Kidney Dis 2001; 37 (1): 1-10<br />

activation potencies of systemic glucocorticoids. Steroids<br />

214. Bicknell AB. Identification of the adrenal protease that cleaves 2001; 66 (7): 597-604<br />

pro-γ-MSH: the dawning of a new era in adrenal physiology? J 232. Santini G, Patrignani P, Sciulli MG, et al. The human pharma-<br />

Endocrinol 2002; 172 (3): 405-10<br />

cology of monocyte cyclooxygenase 2 inhibition by cortisol<br />

© 2005 Adis Data Information BV. All rights reserved. Clin Pharmacokinet 2005; 44 (1)


96 Czock et al.<br />

and synthetic glucocorticoids. Clin Pharmacol Ther 2001; 70 case of aminoglycosides. Clin Pharmacokinet 2000; 38 (4):<br />

(5): 475-83 367-75<br />

233. Jaffuel D, Demoly P, Gougat C, et al. Transcriptional potencies 250. Ferron GM, Jusko WJ. Species- and gender-related differences<br />

of inhaled glucocorticoids. Am J Respir Crit Care Med 2000; in cyclosporine/prednisolone/sirolimus interactions in whole<br />

162 (1): 57-63<br />

blood lymphocyte proliferation assays. J Pharmacol Exp Ther<br />

234. Hogger P, Erpenstein U, Rohdewald P, et al. Biochemical 1998; 286 (1): 191-200<br />

characterization of a glucocorticoid-induced membrane prosteroid<br />

251. Markham A, Faulds D. Theophylline: a review of its potential<br />

tein (RM3/1) in human monocytes and its application as model<br />

sparing effects in asthma. Drugs 1998; 56 (6): 1081-91<br />

system for ranking glucocorticoid potency. Pharm Res 1998; 252. Ito K, Lim S, Caramori G, et al. A molecular mechanism of<br />

15 (2): 296-302<br />

action of theophylline: induction of histone deacetylase ac-<br />

235. Oehling AG, Akdis CA, Schapowal A, et al. Suppression of the tivity to decrease inflammatory gene expression. Proc Natl<br />

immune system by oral glucocorticoid therapy in bronchial Acad Sci USA 2002; 99 (13): 8921-6<br />

asthma. Allergy 1997; 52 (2): 144-54<br />

253. Ito K, Lim S, Caramori G, et al. Cigarette smoking reduces<br />

236. Ducloux D, Carron PL, Motte G, et al. Lymphocyte subsets and histone deacetylase 2 expression, enhances cytokine expresassessment<br />

of cancer risk in renal transplant recipients. Transpl sion, and inhibits glucocorticoid actions in alveolar macro-<br />

Int 2002; 15 (8): 393-6<br />

phages. FASEB J 2001; 15 (6): 1110-2<br />

237. Ducloux D, Carron PL, Rebibou JM, et al. CD4 lymphoglucocorticoid<br />

responsiveness in patients with asthma: results<br />

254. Spahn JD, Fost DA, Covar R, et al. Clarithromycin potentiates<br />

cytopenia as a risk factor for skin cancers in renal transplant<br />

recipients. Transplantation 1998; 65 (9): 1270-2<br />

of a pilot study. Ann Allergy Asthma Immunol 2001; 87 (6):<br />

501-5<br />

238. Ducloux D, Challier B, Saas P, et al. CD4 cell lymphopenia and<br />

atherosclerosis in renal transplant recipients. J Am Soc Neglucocorticoids<br />

and β2 agonists on bronchial airway smooth<br />

255. Roth M, Johnson PR, Rudiger JJ, et al. Interaction between<br />

phrol 2003; 14 (3): 767-72<br />

muscle cells through synchronised cellular signalling. Lancet<br />

239. Chakraborty A, Jusko WJ. Pharmacodynamic interaction of<br />

2002; 360 (9342): 1293-9<br />

recombinant human interleukin-10 and prednisolone using in<br />

256. Boxenbaum H. Pharmacokinetics: philosophy of modeling.<br />

vitro whole blood lymphocyte proliferation. J Pharm Sci 2002;<br />

Drug Metab Rev 1992; 24 (1): 89-120<br />

91 (5): 1334-42<br />

257. DeStefano III JJ, Landaw EM. Multiexponential, multicompart-<br />

240. Ramakrishnan R, Jusko WJ. Interactions of aspirin and salicylic<br />

mental, and noncompartmental modeling. I: methodological<br />

acid with prednisolone for inhibition of lymphocyte proliferalimitations<br />

and physiological interpretations. Am J Physiol<br />

tion. Int Immunopharmacol 2001; 1 (11): 2035-42<br />

1984; 246 (5 Pt 2): R651-64<br />

241. Singh GJ, Adams WP, Lesko LJ, et al. Development of in vivo<br />

258. Meibohm B, Hochhaus G, Mollmann H, et al. A pharmacokinebioequivalence<br />

methodology for dermatologic corticosteroids<br />

tic/pharmacodynamic approach to predict the cumulative corbased<br />

on pharmacodynamic modeling. Clin Pharmacol Ther<br />

tisol suppression of inhaled corticosteroids. J Pharmacokinet<br />

1999; 66 (4): 346-57<br />

Biopharm 1999; 27 (2): 127-47<br />

242. Kelly HW. Establishing a therapeutic index for the inhaled 259. Nichols AI, Jusko WJ. Receptor-mediated prednisolone<br />

corticosteroids (Pt I): pharmacokinetic/pharmacodynamic pharmacodynamics in rats: model verification using a dosecomparison<br />

of the inhaled corticosteroids. J Allergy Clin sparing regimen. J Pharmacokinet Biopharm 1990; 18 (3):<br />

Immunol 1998; 102 (4 Pt 2): S36-51 189-208<br />

243. Meikle AW, Tyler FH. Potency and duration of action of 260. Chan PL, Holford NH. Drug treatment effects on disease proglucocorticoids:<br />

effects of hydrocortisone, prednisone and de- gression. Annu Rev Pharmacol Toxicol 2001; 41: 625-59<br />

xamethasone on human pituitary-adrenal function. Am J Med 261. Dayneka NL, Garg V, Jusko WJ. Comparison of four basic<br />

1977; 63 (2): 200-7 models of indirect pharmacodynamic responses. J Pharma-<br />

244. Yudt MR, Cidlowski JA. The glucocorticoid receptor: coding a cokinet Biopharm 1993; 21 (4): 457-78<br />

diversity of proteins and responses through a single gene. Mol 262. Jusko WJ, Ko HC, Ebling WF. Convergence of direct and<br />

Endocrinol 2002; 16 (8): 1719-26<br />

indirect pharmacodynamic response models. J Pharmacokinet<br />

245. Ito K, Caramori G, Lim S, et al. Expression and activity of Biopharm 1995; 23 (1): 5-8<br />

histone deacetylases in human asthmatic airways. Am J Respir 263. Levy G. Mechanism-based pharmacodynamic modeling. Clin<br />

Crit Care Med 2002; 166 (3): 392-6 Pharmacol Ther 1994; 56 (4): 356-8<br />

246. Hochhaus G, Mollmann H, Derendorf H, et al. Pharmacokine- 264. Sun YN, Jusko WJ. Transit compartments versus gamma distritic/pharmacodynamic<br />

aspects of aerosol therapy using gluco- bution function to model signal transduction processes in<br />

corticoids as a model. J Clin Pharmacol 1997; 37 (10): 881-92 pharmacodynamics. J Pharm Sci 1998; 87 (6): 732-7<br />

247. Mager DE, Moledina N, Jusko WJ. Relative immunosuppres- 265. Nielsen HK, Charles P, Mosekilde L. The effect of single oral<br />

sive potency of therapeutic corticosteroids measured by whole doses of prednisone on the circadian rhythm of serum osteoblood<br />

lymphocyte proliferation. J Pharm Sci 2003; 92 (7): calcin in normal subjects. J Clin Endocrinol Metab 1988; 67<br />

1521-5 (5): 1025-30<br />

248. Kuhnke A, Burmester GR, Krauss S, et al. Bioenergetics of 266. Chakraborty A, Krzyzanski W, Jusko WJ. Mathematical modelimmune<br />

cells to assess rheumatic disease activity and efficacy ing of circadian cortisol concentrations using indirect response<br />

of glucocorticoid treatment. Ann Rheum Dis 2003; 62 (2): models: comparison of several methods. J Pharmacokinet Bio-<br />

133-9 pharm 1999; 27 (1): 23-43<br />

249. Czock D, Giehl M, Keller F. A concept for pharmacokinetic- 267. Rohatagi S, Tauber U, Richter K, et al. Pharmacokinetic/<br />

pharmacodynamic dosage adjustment in renal impairment: the pharmacodynamic modeling of cortisol suppression after oral<br />

© 2005 Adis Data Information BV. All rights reserved. Clin Pharmacokinet 2005; 44 (1)


Pharmacokinetics/Pharmacodynamics of Glucocorticoids 97<br />

administration of fluocortolone. J Clin Pharmacol 1996; 36 with bone marrow transplantation. University of Nebraska<br />

(4): 311-4 Medical Center Bone Marrow Transplant Group. Am J Med<br />

268. Rohatagi S, Bye A, Mackie AE, et al. Mathematical modeling 1994; 96 (4): 327-34<br />

of cortisol circadian rhythm and cortisol suppression. Eur J 286. Bracken MB. Steroids for acute spinal cord injury. Cochrane<br />

Pharm Sci 1996; 4 (6): 341-50<br />

Database Syst Rev 2002; (3): CD001046<br />

269. Ramakrishnan R, DuBois DC, Almon RR, et al. Fifth-genera- 287. de Gans J, van de Beek D. Dexamethasone in adults with<br />

tion model for corticosteroid pharmacodynamics: application bacterial meningitis. N Engl J Med 2002; 347 (20): 1549-56<br />

to steady-state receptor down-regulation and enzyme induction 288. Balis FM, Lester CM, Chrousos GP, et al. Differences in<br />

patterns during seven-day continuous infusion of methyl- cerebrospinal fluid penetration of corticosteroids: possible reprednisolone<br />

in rats. J Pharmacokinet Pharmacodyn 2002; 29 lationship to the prevention of meningeal leukemia. J Clin<br />

(1): 1-24 Oncol 1987; 5 (2): 202-7<br />

270. Hirano T, Oka K, Takeuchi H, et al. A comparison of predniso- 289. Chatham WW, Kimberly RP. Treatment of lupus with corticolone<br />

and methylprednisolone for renal transplantation. Clin steroids. Lupus 2001; 10 (3): 140-7<br />

Transplant 2000; 14 (4 Pt 1): 323-8<br />

290. Illei GG, Austin HA, Crane M, et al. Combination therapy with<br />

271. Tornatore KM, Logue G, Venuto RC, et al. Cortisol pharmaco- pulse cyclophosphamide plus pulse methylprednisolone imdynamics<br />

after methylprednisolone administration in young proves long-term renal outcome without adding toxicity in<br />

and elderly males. J Clin Pharmacol 1997; 37 (4): 304-11 patients with lupus nephritis. Ann Intern Med 2001; 135 (4):<br />

272. Stuck AE, Frey BM, Frey FJ. Kinetics of prednisolone and 248-57<br />

endogenous cortisol suppression in the elderly. Clin Pharmacol 291. Bellomio V, Spindler A, Lucero E, et al. Systemic lupus<br />

Ther 1988; 43 (4): 354-62<br />

erythematosus: mortality and survival in argentina: a multicen-<br />

273. Sharma A, Ebling WF, Jusko WJ. Precursor-dependent indirect ter study. Lupus 2000; 9 (5): 377-81<br />

pharmacodynamic response model for tolerance and rebound 292. Cathcart ES, Idelson BA, Scheinberg MA, et al. Beneficial<br />

phenomena. J Pharm Sci 1998; 87 (12): 1577-84<br />

effects of methylprednisolone “pulse” therapy in diffuse<br />

274. Buttgereit F, da Silva JA, Boers M, et al. Standardised nomen- proliferative lupus nephritis. Lancet 1976; I (7952): 163-6<br />

clature for glucocorticoid dosages and glucocorticoid treat- 293. Honma M, Ichikawa Y, Akizuki M, et al. Double blind trial of<br />

ment regimens: current questions and tentative answers in pulse methylprednisolone versus conventional oral prednisorheumatology.<br />

Ann Rheum Dis 2002; 61 (8): 718-22 lone in lupus nephritis [in Japanese]. Ryumachi 1994; 34 (3):<br />

275. Hayball PJ, Cosh DG, Ahern MJ, et al. High dose oral methyl- 616-27<br />

prednisolone in patients with rheumatoid arthritis: pharmaco- 294. Barron KS, Person DA, Brewer Jr EJ, et al. Pulse methylkinetics<br />

and clinical response. Eur J Clin Pharmacol 1992; 42 prednisolone therapy in diffuse proliferative lupus nephritis. J<br />

(1): 85-8 Pediatr 1982; 101 (1): 137-41<br />

276. Needs CJ, Smith M, Boutagy J, et al. Comparison of methyl- 295. Rose GM, Cole BR, Robson AM. The treatment of severe<br />

prednisolone (1g IV) with prednisolone (1g orally) in rheuma- glomerulopathies in children using high dose intravenous<br />

toid arthritis: a pharmacokinetic and clinical study. J Rheu- methylprednisolone pulses. Am J Kidney Dis 1981; 1 (3): 148-<br />

matol 1988; 15: 224-8 56<br />

277. Feduska NJ, Turcotte JG, Gikas PW, et al. Reversal of renal 296. Eyanson S, Passo MH, Aldo-Benson MA, et al. Methylallograft<br />

rejection with intravenous methylprednisolone prednisolone pulse therapy for nonrenal lupus erythematosus.<br />

“pulse” therapy. J Surg Res 1972; 12 (3): 208-15 Ann Rheum Dis 1980; 39 (4): 377-80<br />

278. August C, Schmid KW, Dietl KH, et al. Prognostic value of 297. Pozzi C, Bolasco PG, Fogazzi GB, et al. Corticosteroids in IgA<br />

lymphocyte apoptosis in acute rejection of renal allografts. nephropathy: a randomised controlled trial. Lancet 1999; 353<br />

Transplantation 1999; 67 (4): 581-5 (9156): 883-7<br />

279. Bolton WK, Sturgill BC. Methylprednisolone therapy for acute 298. Murnaghan K, Vasmant D, Bensman A. Pulse methylpredcrescentic<br />

rapidly progressive glomerulonephritis. Am J Ne- nisolone therapy in severe idiopathic childhood nephrotic synphrol<br />

1989; 9 (5): 368-75 drome. Acta Paediatr Scand 1984; 73 (6): 733-9<br />

280. O’Neill Jr WM, Etheridge WB, Bloomer HA. High-dose corti- 299. Gertz MA, Garton JP, Greipp PR, et al. A phase II study of<br />

costeroids: their use in treating idiopathic rapidly progressive high-dose methylprednisolone in refractory or relapsed multiglomerulonephritis.<br />

Arch Intern Med 1979; 139 (5): 514-8 ple myeloma. Leukemia 1995; 9 (12): 2115-8<br />

281. Vassalotti JA, Delgado FA, Whelton A. Atheroembolic renal 300. Wong P, Cuello C, Bertouch JV, et al. The effects of pulse<br />

disease. Am J Ther 1996; 3 (7): 544-9<br />

methylprednisolone on matrix metalloproteinase and tissue<br />

282. Santostefano M, Cocchi R, Fabbri A, et al. Steroid treatment of inhibitor of metalloproteinase-1 expression in rheumatoid aratheroembolic<br />

renal disease [abstract]. Nephrol Dial Trans- thritis. Rheumatology (Oxf) 2000; 39 (10): 1067-73<br />

plant 2003; 18 Suppl. 4: 622<br />

301. Kirwan JR. The effect of glucocorticoids on joint destruction in<br />

283. Barile LA, Jara LJ, Medina-Rodriguez F, et al. Pulmonary rheumatoid arthritis: the arthritis and rheumatism council lowhemorrhage<br />

in systemic lupus erythematosus. Lupus 1997; 6 dose glucocorticoid study group. N Engl J Med 1995; 333 (3):<br />

(5): 445-8 142-6<br />

284. Leatherman JW, Davies SF, Hoidal JR. Alveolar hemorrhage 302. Engel T, Dirksen A, Frolund L, et al. Methylprednisolone pulse<br />

syndromes: diffuse microvascular lung hemorrhage in immune therapy in acute severe asthma: a randomized, double-blind<br />

and idiopathic disorders. Medicine (Baltimore) 1984; 63 (6): study. Allergy 1990; 45 (3): 224-30<br />

343-61 303. Bowler SD, Mitchell CA, Armstrong JG. Corticosteroids in<br />

285. Metcalf JP, Rennard SI, Reed EC, et al. Corticosteroids as acute severe asthma: effectiveness of low doses. Thorax 1992;<br />

adjunctive therapy for diffuse alveolar hemorrhage associated 47 (8): 584-7<br />

© 2005 Adis Data Information BV. All rights reserved. Clin Pharmacokinet 2005; 44 (1)


98 Czock et al.<br />

304. Greos LS, Vichyanond P, Bloedow DC, et al. Methylpred- requiring a high-dose inhaled steroid, budesonide, to control<br />

nisolone achieves greater concentrations in the lung than pred- moderate to severe asthma. Am Rev Respir Dis 1989; 140 (3):<br />

nisolone: a pharmacokinetic analysis. Am Rev Respir Dis 624-8<br />

1991; 144 (3 Pt 1): 586-92<br />

315. Keller F, Hemmen T, Schoneshofer M, et al. Pharmacokinetics<br />

305. MacLaren R, Jung R. Stress-dose corticosteroid therapy for of methylprednisolone and rejection episodes in kidney transsepsis<br />

and acute lung injury or acute respiratory distress syn- plant patients. Transplantation 1995; 60 (4): 330-3<br />

drome in critically ill adults. Pharmacotherapy 2002; 22 (9):<br />

1140-56<br />

316. Briggs WA, Gimenez LF, Samaniego-Picota M, et al. Relationship<br />

between lymphocyte and clinical steroid responsiveness<br />

306. Luce JM, Montgomery AB, Marks JD, et al. Ineffectiveness of<br />

in focal segmental glomerulosclerosis. J Clin Pharmacol 2000;<br />

high-dose methylprednisolone in preventing parenchymal lung<br />

40 (2): 115-23<br />

injury and improving mortality in patients with septic shock.<br />

Am Rev Respir Dis 1988; 138 (1): 62-8<br />

317. Pollak R, Dumble LJ, Lazda VA, et al. Utility of an in vitro<br />

307. Bernard GR, Luce JM, Sprung CL, et al. High-dose corticosterplant<br />

Proc 1991; 23 (1 Pt 2): 1113-4<br />

immunoassay to guide immunosuppressive therapy. Transoids<br />

in patients with the adult respiratory distress syndrome. N<br />

Engl J Med 1987; 317 (25): 1565-70<br />

318. Langhoff E, Ladefoged J. The impact of high lymphocyte<br />

308. Annane D, Sebille V, Charpentier C, et al. Effect of treatment sensitivity to glucocorticoids on kidney graft survival in pawith<br />

low doses of hydrocortisone and fludrocortisone on mortion<br />

tients treated with azathioprine and cyclosporine. Transplanta-<br />

tality in patients with septic shock. JAMA 2002; 288 (7): 862-<br />

1987; 43 (3): 380-4<br />

71<br />

319. Poznansky MC, Gordon AC, Douglas JG, et al. Resistance to<br />

309. Meduri GU, Headley AS, Golden E, et al. Effect of prolonged methylprednisolone in cultures of blood mononuclear cells<br />

methylprednisolone therapy in unresolving acute respiratory from glucocorticoid-resistant asthmatic patients. Clin Sci<br />

distress syndrome: a randomized controlled trial. JAMA 1998; (Lond) 1984; 67 (6): 639-45<br />

280 (2): 159-65<br />

320. Kirkham BW, Corkill MM, Davison SC, et al. Response to<br />

310. van Vollenhoven RF. Corticosteroids in rheumatic disease: un- glucocorticoid treatment in rheumatoid arthritis: in vitro cell<br />

derstanding their effects is key to their use. Postgrad Med mediated immune assay predicts in vivo responses. J Rheu-<br />

1998; 103 (2): 137-42 matol 1991; 18 (6): 821-5<br />

311. Curtis JJ, Galla JH, Woodford SY, et al. Comparison of daily<br />

321. Liddle GW. Clinical pharmacology of the anti-inflammatory<br />

and alternate-day prednisone during chronic maintenance thersteroids.<br />

Clin Pharmacol Ther 1961; 2 (5): 615-35<br />

apy: a controlled crossover study. Am J Kidney Dis 1981; 1<br />

(3): 166-71 322. Zellner D, Frankewitsch T, Simon S, et al. Statistical analysis of<br />

312. Hunder GG, Sheps SG, Allen GL, et al. Daily and alternate-day<br />

heterogeneous pharmacokinetic data from the literature. Eur J<br />

corticosteroid regimens in treatment of giant cell arteritis:<br />

Clin Chem Clin Biochem 1996; 34 (7): 585-9<br />

comparison in a prospective study. Ann Intern Med 1975; 82<br />

(5): 613-8<br />

313. Groves RW, Toms GC, Houghton BJ, et al. Corticosteroid<br />

Correspondence and offprints: Dr Frieder Keller, Departreplacement<br />

therapy: twice or thrice daily? J R Soc Med 1988; ment of Internal Medicine, Division of Nephrology,<br />

81 (9): 514-6 University Hospital Ulm, Robert-Koch-Str. 8, Ulm 89081,<br />

314. Malo JL, Cartier A, Merland N, et al. Four-times-a-day dosing Germany.<br />

frequency is better than a twice-a-day regimen in subjects E-mail: frieder.keller@medizin.uni-ulm.de<br />

© 2005 Adis Data Information BV. All rights reserved. Clin Pharmacokinet 2005; 44 (1)

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!