Vous êtes sur la page 1sur 21

JPT-06668; No of Pages 21

Pharmacology & Therapeutics xxx (2014) xxx–xxx

Contents lists available at ScienceDirect

Pharmacology & Therapeutics


journal homepage: www.elsevier.com/locate/pharmthera

Azithromycin: Mechanisms of action and their relevance for


clinical applications
Michael J. Parnham a,b,c,⁎, Vesna Erakovic Haber d, Evangelos J. Giamarellos-Bourboulis e,f, Gianpaolo Perletti g,h,
Geert M. Verleden i, Robin Vos i
a
Fraunhofer Institute for Molecular Biology and Applied Ecology, Project Group Translational Medicine and Pharmacology, Frankfurt am Main, Germany
b
Institute of Pharmacology for Life Scientists, Goethe University Frankfurt, Frankfurt am Main, Germany
c
Institute of Clinical Pharmacology, Goethe University Frankfurt, Frankfurt am Main, Germany
d
Fidelta d.o.o., Zagreb, Croatia
e
4th Department of Internal Medicine, University of Athens, Medical School, Athens, Greece
f
Integrated Research and Treatment Center, Center for Sepsis Control and Care, Jena University Hospital, Jena, Germany
g
Biomedical Research Division, Department of Theoretical and Applied Sciences, University of Insubria, Busto A., Varese, Italy
h
Department of Basic Medical Sciences, Ghent University, Ghent, Belgium
i
Respiratory Division, Lung Transplantation Unit, University Hospitals Leuven and Department of Clinical and Experimental Medicine, KU Leuven, Belgium

a r t i c l e i n f o a b s t r a c t

Keywords: Azithromycin is a macrolide antibiotic which inhibits bacterial protein synthesis, quorum-sensing and reduces
Azithromycin the formation of biofilm. Accumulating effectively in cells, particularly phagocytes, it is delivered in high
Macrolide antibiotic concentrations to sites of infection, as reflected in rapid plasma clearance and extensive tissue distribution.
Mechanisms of action Azithromycin is indicated for respiratory, urogenital, dermal and other bacterial infections, and exerts
Pharmacokinetics immunomodulatory effects in chronic inflammatory disorders, including diffuse panbronchiolitis, post-
Immunomodulation
transplant bronchiolitis and rosacea. Modulation of host responses facilitates its long-term therapeutic benefit
Clinical efficacy
in cystic fibrosis, non-cystic fibrosis bronchiectasis, exacerbations of chronic obstructive pulmonary disease
(COPD) and non-eosinophilic asthma.
Initial, stimulatory effects of azithromycin on immune and epithelial cells, involving interactions with
phospholipids and Erk1/2, are followed by later modulation of transcription factors AP-1, NFκB, inflammatory
cytokine and mucin release. Delayed inhibitory effects on cell function and high lysosomal accumulation
accompany disruption of protein and intracellular lipid transport, regulation of surface receptor expression, of
macrophage phenotype and autophagy. These later changes underlie many immunomodulatory effects of
azithromycin, contributing to resolution of acute infections and reduction of exacerbations in chronic airway
diseases. A sub-group of post-transplant bronchiolitis patients appears to be sensitive to azithromycin, as may
be patients with severe sepsis. Other promising indications include chronic prostatitis and periodontitis, but
weak activity in malaria is unlikely to prove crucial. Long-term administration of azithromycin must be balanced
against the potential for increased bacterial resistance. Azithromycin has a very good record of safety, but recent
reports indicate rare cases of cardiac torsades des pointes in patients at risk.
© 2014 Elsevier Inc. All rights reserved.

Abbreviations: AECOPD, acute exacerbations of COPD; AP-1, activator protein-1; ARDS, acute respiratory distress syndrome; AUC, area under the plasma concentration-versus-time
curve; BAL, broncho-alveolar lavage; BOS, bronchiolitis obliterans syndrome; CAD, cationic amphiphilic drugs; CAP, community acquired pneumonia; CBP, chronic bacterial prostatitis;
CF, cystic fibrosis; CFTR, CF transmembrane conduction regulatory protein; Cmax, peak plasma concentration; COPD, chronic obstructive pulmonary disease; COX, cyclo-oxygenase;
cPLA2, cytoplasmic phospholipase A2; DNFB, dinitrofluorobenzene; DPB, diffuse panbronchiolitis; DTH, delayed type hypersensitivity; ERK, extracellular signal-regulated kinase; FGF, fi-
broblast growth factor; GM-CSF, granulocyte–macrophage colony stimulating factor; GVH, graft versus host reaction; HR, hazard ratio; IL, interleukin; i.n, intranasal; i.p, intraperitoneal; i.t,
intratracheal; i.v, intravenous; JNK, c-Jun NH(2)-terminal kinase; LC3, microtubule-associated protein 1A/light chain 3; LOS, length of stay; LPS, bacterial lipopolysaccharide; LRTI, lower
respiratory tract infection; MAPK, mitogen-activated protein kinase; MDR1, multidrug resistance protein 1; MIC, minimum inhibitory concentration; MLSbK, macrolides, lincosamines,
streptogramin B and ketolides; MMP, metalloproteinase; MODS, multiple organ dysfunction syndrome; MPO, myeloperoxidase; MUC5AC, mucin 5AC; MV, mechanical ventilation;
NFκB, nuclear factor kappaB; NGU, non-gonococcal urethritis; NK, natural killer cells; NR, not reported; OR, odds ratio; p.o, oral; PG, prostaglandin; PI, phosphatidylinositol; PI3K,
phosphoinositide-3-kinase; PK, pharmacokinetic; PMNL, polymorphonuclear leukocytes; PS, phosphatidylserine; RCT, randomized controlled clinical trial; ROS, reactive oxygen species;
SAA, serum amyloid A protein; SMC, smooth muscle cell; SREBP, sterol regulatory element binding protein; STAT, signal transducers and activators of transcription; STD, sexually trans-
mitted disease; t1/2, plasma half-life; TLR, Toll-like receptor; Tmax, time to Cmax; TNFα, tumor necrosis factor alpha; Th, T helper cell; VAP, ventilation assisted pneumonia; Vd, volume of
distribution; VEGF, vascular endothelial growth factor.
⁎ Corresponding author at: Fraunhofer IME-TMP, Theodor-Stern-Kai 7, D-60596 Frankfurt am Main, Germany. Tel.: +49 69 6301 84234.
E-mail addresses: michael.parnham@ime.fraunhofer.de (M.J. Parnham), Vesna.ErakovicHaber@glpg.com (V.E. Haber), egiamarel@med.uoa.gr (E.J. Giamarellos-Bourboulis),
gianpaolo.perletti@uninsubria.it (G. Perletti), geert.verleden@uzleuven.be (G.M. Verleden), robin.vos@uzleuven.be (R. Vos).

http://dx.doi.org/10.1016/j.pharmthera.2014.03.003
0163-7258/© 2014 Elsevier Inc. All rights reserved.

Please cite this article as: Parnham, M.J., et al., Azithromycin: Mechanisms of action and their relevance for clinical applications, Pharmacology &
Therapeutics (2014), http://dx.doi.org/10.1016/j.pharmthera.2014.03.003
2 M.J. Parnham et al. / Pharmacology & Therapeutics xxx (2014) xxx–xxx

Contents

1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 0
2. Antibacterial mechanisms of action . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 0
3. Pharmacokinetics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 0
4. Immunomodulatory activities . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 0
5. Clinical applications in respiratory infections and airway inflammation . . . . . . . . . . . . . . . . . . . . . 0
6. Infections of the urogenital tract and sexually-transmitted infections . . . . . . . . . . . . . . . . . . . . . . 0
7. Clinical application in sepsis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 0
8. Other clinical applications . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 0
9. Safety . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 0
10. Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 0
Conflict of interest . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 0
Acknowledgments . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 0
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 0

1. Introduction properties and immunomodulatory actions contribute to the effects


observed.
Azithromycin, a second generation macrolide, broad-spectrum
antibacterial, has received increasing attention in recent years because 2. Antibacterial mechanisms of action
of additional effects on host-defence reactions and chronic human
diseases. It is the prototype 15-membered lactone ring azalide, synthe- 2.1. Inhibitory actions
sized in the early 1980s as a semi-synthetic derivative of erythromycin.
Discovered around the same time by researchers at Pfizer in the United Azithromycin, like other macrolide antibiotics, inhibits bacterial
States (Bright & Hauske, 1984) and at PLIVA, Croatia (Kobrehel et al., protein synthesis by binding to and interfering with the assembly of
1982), PLIVA patented first and licensed the compound to Pfizer. the 50S large ribosomal subunit and the growth of the nascent poly-
With much improved pharmacokinetic properties over erythromycin, peptide chain (Champney & Burdine, 1998; Champney et al., 1998;
azithromycin became the most widely used broad-spectrum antibacte- Hansen et al., 2002). It binds at the polypeptide exit tunnel, close to
rial in North America. Pfizer Inc's Arthur E. Girard and Gene Michael the peptidyl transferase center (PTC) on the 23S rRNA, but does not
Bright, together with PLIVA's Slobodan Djokic (posthumously) and inhibit PT activity, in contrast to larger macrocyclic antibiotics. The
Gabrijela Kobrehel, received in 2000 the American Chemical Society's basicity of azithromycin leads to faster penetration of the outer mem-
award of “Heroes of Chemistry who have promoted human welfare in branes and more effective entrance into the bacteria, thereby enhancing
the area of health” for their discovery of Zithromax® (azithromycin). activity against Gram-negative bacteria (Dinos et al., 2001).
Azithromycin shares the same mechanism of antibacterial action Binding sites on the bacterial ribosome for the structurally different
as other macrolide antibiotics (Allen, 2002), but accumulates more macrolides, lincosamines, streptogramin B and ketolides (MLSbK) over-
effectively in phagocytes, thus being delivered in high concentrations lap significantly, so that changes in a single ribosomal region simulta-
to sites of infection (Miossec-Bartoli et al., 1999; Wilms et al., 2006). neously alter susceptibility to various MLSbK antibiotics.
It also inhibits bacterial quorum-sensing and reduces formation of Although ineffective as a bactericidal agent against Pseudomonas
biofilm and mucus production, which extend its range of antibacterial aeruginosa at clinically relevant concentrations, azithromycin inhibits
actions (Tateda et al., 2001; Hoffmann et al., 2007). As an antibiotic, the generation of both growth-stimulating, quorum-sensing com-
azithromycin is indicated for respiratory, urogenital, dermal and other pounds, and alginate biofilm which protects the micro-organism from
bacterial infections, but has beneficial effects in chronic inflammatory antibiotic actions (Tateda et al., 2001; Hoffmann et al., 2007). Efficacy
disorders such as diffuse panbronchiolits, bronchiolitis obliterans and against both P. aeruginosa virulence factor production and biofilm
rosacea. Efficacy in these conditions is ascribed to immunomodulatory formation, as well as its ability to decrease the minimum inhibitory
effects on innate and adaptive immune responses. Modulation of host concentration (MIC) of anti-pseudomonas agents (Lutz et al., 2012)
response reactions also accounts, at least partially, for beneficial effects are related to ribosomal protein synthesis inhibition (Kohler et al.,
in cystic fibrosis, non-cystic fibrosis bronchiectasis, bronchial obliterans 2007).
syndrome (BOS) and chronic obstructive pulmonary disease (COPD). Azithromycin also shows moderate activity against the malaria par-
Azithromycin is well-tolerated and has a very good record of safety. asite, Plasmodium spp., exhibited as delayed killing, which is achieved
Although macrolides have a class warning for potential cardiac QT pro- via the apicoplast (Dahl & Rosenthal, 2007). Clinical studies, however,
longation, azithromycin does not show this effect under experimental failed to demonstrate even equivalence of three-day treatment with
conditions (Milberg et al., 2002). Until recently, only a handful of cases azithromycin to other antimalarials or drug combinations (van Eijk &
of QT prolongation had been reported for patients treated with the Terlouw, 2011).
drug (Kezerashvili et al., 2007). This is mainly because azithromycin,
unlike other macrolide antibiotics, does not interact with CYP3A4, 2.2. Resistance phenotypes
despite a minor interaction with the anti-coagulant warfarin (Kanoh
& Rubin, 2010; Mergenhagen et al., 2013). Recently, evidence for Bacteria resist azithromycin in two ways: a) by changing the target/
increased risk of QT prolongation with azithromycin has appeared, but binding site via methylation of key rRNA nucleotides or mutation of
mainly in patients with greater susceptibility to adverse cardiac effects some ribosomal components and b) by efflux pump activity, thereby
(Giudicessi & Ackerman, 2013). decreasing its intrabacterial accumulation.
We review here the antibacterial actions and pharmacokinetics of Ribosome methylation is the most important resistance mechanism
azithromycin, as well as its immunomodulatory effects and the for all MLSbK antibiotics (Sutcliffe & Leclerq, 2002). The MLSbK-II
mechanisms involved, in comparison to other macrolide antibiotics. phenotype, characterized by the presence of Erm methylases, is highly
We discuss the main clinical uses of azithromycin, drawing attention resistant to all MLSbK antibiotics, including azithromycin, which
to emerging indications and emphasising how its pharmacokinetic induces erm genes in Streptococci and in Staphylococci (Sutcliffe &

Please cite this article as: Parnham, M.J., et al., Azithromycin: Mechanisms of action and their relevance for clinical applications, Pharmacology &
Therapeutics (2014), http://dx.doi.org/10.1016/j.pharmthera.2014.03.003
M.J. Parnham et al. / Pharmacology & Therapeutics xxx (2014) xxx–xxx 3

Leclerq, 2002). In S. pyogenes there are three types of inducible Falkner, 1990) as a result of high cellular accumulation, particularly in
phenotype (a) iMLSbK-A, (b) iMLSbK-B, which is highly resistant to phagocytes, in which at least 200-fold higher intracellular than extracel-
azithromycin (MICN128 μg/mL) and (c) iMLSbK-C with a medium resis- lular concentrations can be achieved (Bosnar et al., 2005; Stepanic et al.,
tance profile (MICs up to 16 μg/mL) against azithromycin (Giovanetti 2011). Consequently, specific accumulation of azithromycin occurs at
et al., 1999; Giovanetti et al., 2002). In addition to methylation of sites of inflammation (Miossec-Bartoli et al., 1999; Wilms et al., 2006).
23rRNA, resistance can be achieved by mutation of various genes coding In drug-free medium, preloaded cells release azithromycin slowly,N80%
for ribosomal components. This type of resistance has been observed being retained, so cellular accumulation and retention account for the
infrequently in S. pneumonia, S. pyogenes and H. influenzae, the most long half-life of azithromycin in vivo (Bosnar et al., 2005; Stepanic
common mutations occurring in domain V rRNA and proteins L4 and et al., 2011). Accumulation of azithromycin in polymorphonuclear
L22 (Tait-Kamradt et al., 2000; Franceschi et al., 2004). cells (PMNLs) acts as an innate targeted delivery system to the site of in-
Mef transporter (efflux pump) expression typically results in a lower fection. Degranulation of PMNs by bacteria then releases azithromycin
level of specific resistance towards 14- and 15-membered macrolide into the extracellular space, where it exerts its antibacterial activity.
antibiotics (MICs 4–32 ug/mL) and this type of resistance is referred Azithromycin also accumulates, to a less and varying extent, in other
to as the M phenotype (Leclercq, 2002). cells such as epithelial cells, fibroblasts, lymphocytes, hepatocytes
Increased clinical use of macrolide antibiotics is linked with an (Bosnar et al., 2005; Gladue & Snider, 1990; Matijasic et al., 2012;
increase in pneumococcal macrolide resistance. The Alexander Project, Stepanic et al., 2011). Fibroblasts, due to their wide distribution, have
monitoring antibiotic resistance among respiratory pathogens, reported been proposed as a potential reservoir for azithromycin, slowly releas-
a global rate of pneumococcal macrolide resistance between 16.5% and ing or passing it to nearby phagocytes for transport to the site of infec-
21.9% for 1996 and 1997 (Felmingham & Gruneberg, 2000), increasing tion (Gladue & Snider, 1990; McDonald & Pruul, 1991). Based on such
to 24.6% from 1998 to 2000 (M. R. Jacobs & Johnson, 2003). Over this retention in epithelial lining and bronchoalveolar cells, macrolides,
period, macrolide-exceeded penicillin-resistance in 19/26 countries. including azithromycin, find veterinary use for the treatment of pneu-
The PROTEKT US surveillance study reported similar findings. From monia (Villarino & Martin-Jimenez, 2013).
2000 to 2004, erythromycin resistance among S. pneumoniae isolates Most intracellular azithromycin is located in acidic organelles such
remained high (Jenkins et al., 2008), mef(A) being the most common as lysosomes. Initially shown to concentrate in the cytoplasm and plas-
macrolide resistance genotype, accounting for N60% of tested isolates malemma of PMNLs and monocytes (Wildfeuer et al., 1993), the vast
(Farrell & Jenkins, 2004). Although macrolide resistance rates have majority of azithromycin in these cells is associated with the granular
subsequently stabilized, the prevalence of clonal isolates with a com- fraction (Carlier et al., 1994; Miossec-Bartoli et al., 1999). Recent data
bined erm(B) and mef(A) genotype and a high-level of macrolide and support its lysosomal localization in hepatocytes (Matijasic et al.,
multidrug resistance has increased. 2012). In keeping with lysosomal accumulation, azithromycin causes
Distribution of macrolide resistance genotypes varies between and lysosomal pH change and functional impairment in murine macro-
within countries. In the studies mentioned above, highest resistance phages (Nujic et al., 2012a).
rates were in Asia and some European countries (France, Italy, Spain, The exact mechanism of cellular accumulation remains unclear. One
Belgium), being N70% and 25–60%, respectively. Macrolide resistance possibility is passive transmembrane transport and trapping by proton-
was common among S. pneumoniae isolates (77.2–81.9%) in Japan ation in acidic organelles, as this can be inhibited simply by neutraliza-
during PROTEKT years 1999–2004, the erm(B) genotype accounting tion of lysosomal pH, markedly increasing azithromycin efflux (Hand &
for most resistance (Inoue et al., 2008). In Portugal (1994–2002), Hand, 2001). Moreover, macrolide cellular accumulation and retention
emergence of macrolide-resistant S. pneumoniae strains clearly correlat- can be predicted by experimentally determined physicochemical
ed with increased azithromycin use (Dias & Canica, 2004). parameters and the calculated number of positively charged atoms
In a recent study on commensal Staphylococcus aureus in Europe, (Stepanic et al., 2011). Like other cationic amphiphilic drugs (CADs),
except for penicillin, the highest recorded resistance rate was to macrolides may degrade cell membranes, together with which the
azithromycin (from 1.6% in Sweden to 16. 9% in France) (den Heijer drugs are carried to lysosomes (Baciu et al., 2006).
et al., 2013). As usage of azithromycin and clarithromycin has increased An active uptake mechanism, mediated by transporter protein(s), is
for indications beyond RTI, resistance has started to emerge in non- supported by dependence of accumulation on cell viability, tempera-
respiratory pathogens. For instance, macrolide treatment of healthy ture, pH and extracellular calcium concentration (Gladue et al., 1989;
volunteers, in comparison to placebo, significantly increased the Mtairag et al., 1995; Pascual et al., 1995). Macrolides also compete for
proportion of macrolide-resistant oral streptococci for up to 180 days cell accumulation (Munic et al., 2010). They are inhibitors of ATP-
after treatment (Malhotra-Kumar et al., 2007). Consequently, as with binding cassette transporters such as P-glycoprotein (multidrug
all antibiotics, azithromycin should be administered for all relevant resistance protein 1, MDR1) (Wang et al., 2000; Yasuda et al., 2002;
conditions with due consideration of the associated benefits and risks. Asakura et al., 2004; Munic et al., 2010), multidrug resistance associated
protein 2 (MRP2) (Sugie et al., 2004) and organic anion-transporting
3. Pharmacokinetics polypeptide, OATP (Garver et al., 2008) and substrates for MDR1.
Ranking of macrolides for cellular accumulation is proportional to
Early pharmacokinetic (PK) studies on azithromycin, indicating low cellular expression of MDR1, its overexpression decreasing macrolide
plasma levels, were incorrectly interpreted as reflecting poor PK proper- accumulation 2- to 80-fold, particularly with azithromycin and erythro-
ties. Azithromycin is now known to have a large volume of distribution, mycin (Munic et al., 2010).
achieving high tissue concentrations and is efficiently delivered to sites It seems likely that both passive and active mechanisms are in-
of infection. Compared with older generation macrolides, it is more volved, variability in macrolide accumulation resulting from changes
stable in acidic media and has a longer half-life, allowing for once a in acidic organelles or in different transporter proteins.
day or even single dose treatment (Amsden, 1995, 2001). Moreover,
azithromycin is barely metabolized, with no active metabolites and
does not induce CYPs (Amsden, 1995; Nightingale, 1997). 3.2. Pharmacokinetic properties

3.1. Cellular accumulation Extensive data are available on the clinical PK of various formula-
tions of azithromycin, both under normal and pathologic conditions
Concentrations of azithromycin achieved in tissues, are up to 100- (Table 1). Knowledge of the salient kinetic features is crucial to an
fold higher than those in plasma (Foulds et al., 1990; Shepard & understanding of its therapeutic effects.

Please cite this article as: Parnham, M.J., et al., Azithromycin: Mechanisms of action and their relevance for clinical applications, Pharmacology &
Therapeutics (2014), http://dx.doi.org/10.1016/j.pharmthera.2014.03.003
4 M.J. Parnham et al. / Pharmacology & Therapeutics xxx (2014) xxx–xxx

Azithromycin shows plasma protein binding of approximately 30% factors than drug metabolism are thought to account for serious adverse
and relatively low oral bioavailability (17–37%) (Foulds et al., 1990; effects seen with azithromycin (Fleet et al., 2013) (see Section 9).
Luke & Foulds, 1997). Following a single standard 500 mg oral dose, PK parameters may be affected considerably by co-administration of
peak plasma concentrations of 0.2–0.4 mg/L are attained, with a Tmax azithromycin with food, prolongation of Tmax, 50% reduction of Cmax,
of 2–4 h that may be delayed significantly in elderly subjects (Coates and of AUC0–72h having been observed in fed, compared to fasting pa-
et al., 1991) (Table 1). Single oral doses of 1.5 g azithromycin attain tients (Table 1) (Curatolo et al., 2011). Azithromycin pharmacokinetics,
peak plasma concentrations as high as 1.46 μg/mL (Amsden, 1996). in keeping with low hepatic metabolism, are not significantly affected
The kinetics of azithromycin are best described by a 3-compartment by mild or moderate hepatic impairment or class-A or -B liver cirrhosis
pharmacokinetic model (Ballow et al., 1998; Pene Dumitrescu et al., (Mazzei et al., 1993), nor by mild or moderate renal insufficiency
2013). (Hoffler et al., 1995; Singlas, 1995). PK parameters are also virtually
When administered as 500 mg extended, rather than immediate unchanged in patients with diabetes (Ernst et al., 2000). Hence, no
release tablets, a threefold higher AUC0–24 and twofold higher Cmax dose modifications of azithromycin appear to be necessary in patients
may be attained (Liu et al., 2011; Lucchi et al., 2008) (Table 2). Plasma suffering from these conditions. However, severe renal impairment
clearance of azithromycin is remarkably high (50 L/h), and the volume noticeably increases Cmax and AUC0–120 (http://www.medicines.org.
of distribution (Vd) is approximately 30 L/kg (Singlas, 1995). Hence, low uk/emc/medicine/26131). Interestingly, in such patients on permanent
plasma concentrations are the consequence of rapid blood-tissue distri- peritoneal dialysis due to oliguric end-stage renal failure (Kent et al.,
bution. The plasma half-life of azithromycin is approximately 70 h or 50 2001), the Cmax of azithromycin increased but Tmax appeared to be in
h following oral or intravenous formulations, respectively (Lode et al., the normal range.
1996). This remarkable pharmacokinetic characteristic of azithromycin
allows infrequent dosing (in bacterial urethritis or cervicitis, single 3.3. Tissue distribution
administration may be sufficient), and ensures optimal patient compli-
ance and sustained exposure to the drug. In fact, the AUC0–24 after a The distribution of azithromycin into inflamed tissues and compart-
single 500 mg oral dose is 3 μg∙h/mL (Matzneller et al., 2013), reaching ments has been extensively investigated. In bronchial washings and
as high as 8–11 μg∙h/mL after intravenous administration of 500 mg lung tissue biopsy specimens (Di Paolo et al., 2002), in lungs, in sinus
(Luke & Foulds, 1997; Rodvold et al., 2003). After a standard dose of mucosal tissue (Ehnhage et al., 2008; Fang et al., 2009), in middle ear
500 mg/day for 3 days, much of the drug is present in blood leukocytes fluids in children with otitis media (Pukander & Rautianen, 1996;
(Matzneller et al., 2013; Pene Dumitrescu et al., 2013). On day 3, blood Scaglione et al., 1999), in the aqueous humor and in conjunctival biopsy
concentration was double that in plasma and Vd to the periphery was specimens (Tabbara et al., 1998), in tissues lining periodontal pockets
2980 L, compared with 439 L in plasma. After 30 days, drug concentra- (Gomi et al., 2007), and in inflammatory skin blisters (Ballow et al.,
tions were 4-fold higher in blood than in plasma. In healthy volunteers, 1998), azithromycin concentrations were always higher than those in
much less drug is present in tissues (Matzneller et al., 2013). Release is plasma. This tissue accumulation offsets the sub-optimal absorption of
likely induced by inflammatory stimuli. the drug. In contrast to many other antibiotics, macrolides penetrate
Azithromycin is mainly eliminated unchanged in the feces by both well into prostatic tissue and seminal fluids. For example, Cmax of
biliary excretion and trans-intestinal secretion. Only about 6% (oral azithromycin and clarithromycin in human prostatic tissue after 500
dose) to 12% (i.v.) is recovered unchanged in the urine (Singlas, or 750 mg oral dosing is 2.54 and 3.83 μg/ml, respectively, about 20-
1995). Using ileostomy fluid sampling and correction for the fraction and 2-fold higher than in plasma (Foulds et al., 1991; Giannopoulos
of i.v. azithromycin recovered in the fluid, approximately 45% of an et al., 2001). This is important with regard to the use of azithromycin
oral dose of 500 mg was unabsorbed after intestinal passage up to the in prostatitis (see Section 6.2).
ileal stoma (Luke & Foulds, 1997). Thus, incomplete absorption, rather In summary, azithromycin concentrates to a considerable extent in
than a first-pass effect, is the major cause of the sub-optimal bioavail- many inflamed tissues and fluids. In this respect, its extensive uptake
ability of azithromycin (Luke & Foulds, 1997). into tissue-infiltrating leukocytes is a significant contributing factor
The inactive descladinose and the active 9a-N-desmethyl metabo- (see Sections 3.1 and 4.1.2).
lites are frequently recovered because of acidic degradation in the gut
(Luke & Foulds, 1997; Raines et al., 1998). Hepatic biotransformation 3.4. Pediatric pharmacokinetics
is marginal (Curatolo et al., 2011). Importantly, azithromycin is neither
metabolized by nor an inhibitor of CYP3A4 and therefore, does not The AUC0–∞ of azithromycin (2 g dose) is reduced in pregnant
interact with many commonly used medications that are oxidatively women (Salman et al., 2010), and plasma half-life is similar in both
metabolized by CYP3A4 (Shakeri-Nejad & Stahlmann, 2006). Other placenta and fetus (70 h), though considerably shorter in amniotic

Table 1
Azithromycin pharmacokinetic data extracted from a sample of studies published between the years 1996 and 2013.

(Luke & Foulds, 1997) (Amsden, 1996) (Rodvold et al., 2003) (Matzneller et al., 2013) (Ballow et al., 1998) (Coates et al., 1991)

Dose/route (mg) 500 IV 500 oral 1500 oral 500 oral 500 IV 500 oral 500 oral 500 oral young 500 mg oral elderly
(18–40 years) (65–85 years)
Cmax (μg/mL) 3.40 0.21 1.46 0.54 1 0.46 0.27 0.41 0.38 μg/mL
tmax (h) 0.7 2.5 – – – 2.92 – 2.50 3.80 h
AUC (μg∙h/mL) 11⁎ 1.27⁎ 13.1⁎⁎ 11.2⁎⁎ 8.2⁎ 3.05⁎ – 2.5⁎ 3 μg∙h/mL⁎
t1/2 (h) 50.2 – – – – – 78.6
Ke (h−1) 0.00138 – – – – – –
Clplasma (L/h) 46.6 – – – – – –
F (%) 17.8 – – – – – –
Vd (L/Kg) – 89.5 L/Kg 109.2 – – –

Cmax, peak plasma concentration; tmax, time- to Cmax; Ke, pharmacokinetic elimination constant; Cl, plasmatic clearance; F, oral bioavailability; Vd, distribution volume.
AUC, area under- (definite integral of-) the plasma concentration-versus-time curve
⁎ AUC0–24.
⁎⁎ AUC0–∞.

Please cite this article as: Parnham, M.J., et al., Azithromycin: Mechanisms of action and their relevance for clinical applications, Pharmacology &
Therapeutics (2014), http://dx.doi.org/10.1016/j.pharmthera.2014.03.003
M.J. Parnham et al. / Pharmacology & Therapeutics xxx (2014) xxx–xxx 5

Table 2
Pharmacokinetic data for azithromycin administered in different oral formulations or in fasting vs. fed patients.

(Lucchi et al., 2008) (Liu et al., 2011) (Muto et al., 2011) (Curatolo et al., 2011)

Dose/route 500 mg immediate 2 g extended 500 mg IR oral 2 g extended 2 g extended 500 mg oral fast dissolving
release oral release oral release oral release oral formula in fasting/fed patients

Cmax (mg/L) 0.39 0.94 0.40 0.72 0.4/0.2


tmax (h) 4 4 2 3.5 2/4
AUC0–24 (μg∙h/mL) 3.1 10 2.67 7.87
Clplasma (L/h) 122 103
Vd (L) 518 1830

fluid (3 h) and umbilical cord serum (12 h) (Ramsey et al., 2003). The association with reduced rhinovirus proliferation and release (Gielen
serum half-life in extremely preterm infants is shorter (58 h) than at et al., 2010). Such activities undoubtedly contribute to therapeutic
full term or in adults. Importantly, preterm infants, treated with benefit in respiratory infections.
10 mg/kg intravenous azithromycin, appear to be overexposed (about In airway smooth muscle cells, azithromycin inhibits interleukin (IL)-
3.2-fold) to the drug compared to older children (age: 0.5–2 years) 17-induced IL-8 and 8-isoprostane release, whereas dexamethasone
(Jacobs et al., 2005; Hassan et al., 2011) and clearance is low (0.18 L/h failed to attenuate the IL-8 production (Vanaudenaerde et al., 2007).
in 1 kg pre-term neonate vs 0.98 L/h/kg in older children). This may On the other hand, azithromycin and dexamethasone (more strongly)
be attributed to the immature biliary excretion pathways in preterm both reduce fibroblast growth factor (FGF)-induced vascular endothelial
neonates (Hassan et al., 2011). growth factor (VEGF) production through interaction with p38 mitogen-
In pediatric subjects (mean 24 months) given an immediate-release activated protein kinase (MAPK)-signaling (Willems-Widyastuti et al.,
formulation of azithromycin (30 mg/kg), plasma concentrations and 2013). Azithromycin also demonstrates an anti-proliferative and auto-
drug exposure were higher than in adults (Liu et al., 2011), necessitating phagic effect (Stamatiou et al., 2009) (Section 4.2.2), and directly relaxes
a dose adjustment (Table 1). When an extended release formulation pre-contracted airway smooth muscle cells (Daenas et al., 2006).
was administered to children, the Cmax was still elevated compared to
adult subjects, but the AUC0–24 was lower (Liu et al., 2011). 4.1.2. Leukocytes
In neutrophils, subsequent to its pronounced accumulation (see
4. Immunomodulatory activities Section 3.1), slow efflux of azithromycin prolongs its effects on the
cells in vitro and in vivo (Bosnar et al., 2005; Culic et al., 2002). This
Early studies on macrolide antibiotics and host defense revealed extensive intracellular retention of azithromycin, for up to 28 days in
inhibitory effects on neutrophils in vitro, on experimental inflammation human neutrophils after standard dosing, leads to initial stimulation
(Culic et al., 2001) and particularly actions on cytokine release (Bartold of neutrophil degranulation and phagocytosis-associated oxidative
et al., 2013). While azithromycin was shown in the late 1980s to inhibit burst, further contributing to its antibacterial activity (Culic et al.,
inflammation and lysosomal enzyme release in arthritic rats (Carevic & 2002). Acute inhibitory effects of azithromycin include down-
Djokic, 1988), subsequent studies were carried out predominantly with regulation of neutrophil chemokine production (IL-8, GRO-α, MPO);
erythromycin, clarithromycin and roxithromycin. All macrolide antibi- late effects comprise attenuation of neutrophil oxidative burst re-
otics inhibit various standard models of inflammation (Ianaro et al., sponses, down-regulation of myeloperoxidase (MPO) production,
2000; Culic et al., 2001; Ivetic Tkalcevic et al., 2012) and more recent increased neutrophil apoptosis and attenuation of chemokine (IL-8)-
studies have sought to clarify immunomodulatory mechanisms in vivo and leukotriene (LT)B4-dependent and chemokine-independent
(Table 3) and at the molecular level. neutrophil chemotactic responses by inhibition of transcription factors,
nuclear factor kappa-B (NFκ-B) and activator protein-1 (AP-1) (Culic
4.1. Cellular functions et al., 2001; Culic et al., 2002; Tamaoki et al., 2004; Tsai & Standiford,
2004). Azithromycin also inhibits prostaglandin (PG)E2-synthesis by
Modulation of host defense by azithromycin and other macrolide suppressing expression of prostaglandin synthetic enzymes (COX-1
antibiotics occurs through interaction with structural cells, such as and COX-2) in both lipopolysaccharide (LPS)-stimulated polymorpho-
epithelial or endothelial cells, smooth muscle cells or fibroblasts, as nuclear leukocytes and monocytes (Miyazaki et al., 2003). In the latter,
well as with leukocytes (macrophages, polymorphonuclear leukocytes azithromycin decreases tumor necrosis factor-α (TNF-α) and granulo-
or neutrophils, mononuclear leukocytes or monocytes, T cells and cyte–macrophage colony stimulating factor (GM-CSF) production
dendritic cells). Immunomodulatory effects on cellular functions of the (Khan et al., 1999).
macrolide antibiotics as a class were reviewed a few years ago (Kanoh In sputum cells (81.7% neutrophils, 12.9% monocytes/macrophages)
& Rubin, 2010). isolated from steroid-naïve patients with COPD, azithromycin and
other macrolides inhibited, in a concentration-dependent manner,
4.1.1. Structural cells the release of a variety of inflammatory cytokines and chemokines
In epithelial cells, azithromycin maintains integrity and the (Marjanovic et al., 2011). The macrolides were more broadly inhibitory
transepithelial resistance in the face of permeability induced by than the corticosteroid, dexamethasone, the PDE4 inhibitor, roflumilast,
virulence factors from P. aeruginosa, in conjunction with rearrangement and the p38 kinase inhibitor, SB203580. Azithromycin was less effective
of tight junction adhesion molecules, including claudins (Asgrimsson than clarithromycin and roxithromycin, but its broad inhibitory activity
et al., 2006; Halldorsson et al., 2010). Increased cell integrity is also may provide corticosteroid-sparing effects in treatment of respiratory
accompanied by reduced mucin secretion (Imamura et al., 2004), inflammation (see Section 5.4). An extensive list of the effects of
altered signal transduction (MAPK pathway) (see Section 4.2.4) and azithromycin on cytokine production, generally, was published recently
attenuated expression of inflammatory (e.g. IL-6 and 8, MMP-1, 2, 9, (Bartold et al., 2013). Azithromycin inhibits neutrophilic inflammation
10 and 13, and GM-CSF), lipid metabolism and cell cycle pathways in vivo induced by LPS given by various routes (Table 3). Lung
(Gielen et al., 2010; Murphy et al., 2008b; Ribeiro et al., 2009). In neutrophilia induced by intranasal LPS is inhibited by azithromycin in
bronchial epithelial cells infected with rhinoviruses, azithromycin association with a reduction in lung GM-CSF and IL-1β, both of which
stimulated interferon and interferon-stimulated gene expression, in were also inhibited by azithromycin in macrophages in vitro (Bosnar

Please cite this article as: Parnham, M.J., et al., Azithromycin: Mechanisms of action and their relevance for clinical applications, Pharmacology &
Therapeutics (2014), http://dx.doi.org/10.1016/j.pharmthera.2014.03.003
6 M.J. Parnham et al. / Pharmacology & Therapeutics xxx (2014) xxx–xxx

et al., 2009). Subsequently, the action on lung neutrophilia was localized complex (MHC) class II molecules in LPS-stimulated dendritic cells,
to the alveolar macrophages (Bosnar et al., 2011). and attenuates Toll-like receptor (TLR)-4 expression, IL-12 production
In alveolar macrophages, azithromycin attenuates LPS-induced and allostimulatory capacity, suggesting potential for modulation of
production and expression of pro-inflammatory cytokines through inhi- allogeneic responses (Iwamoto et al., 2011, 2013). Animal studies
bition of AP-1 (Bosnar et al., 2011; Meyer et al., 2009) (see Section 4.2.4) suggest that azithromycin may also modulate Th2 cell responses
and it inhibits arachidonate release and metabolism in LPS-stimulated in vivo, probably by effects on antigen-presenting cells (Table 3). In-
macrophages (Banjanac et al., 2012). Azithromycin increases phago- duction of a modulatory DC phenotype could explain findings in healthy
cytosis, probably by upregulating the macrophage mannose receptor, human volunteers that azithromycin enhanced proliferative and IL-2
CD206 (Hodge et al., 2008) and makes lysosomes in macrophages receptor responses of blood mononuclear cells (MNC) to pokeweed
more resistant to oxidant challenge (Persson et al., 2012). Azithromycin mitogen (Tomazic et al., 1993). An inhibitory effect on antigen-
also attenuates T helper-1 cell (Th-1) responses following LPS or presentation by DCs almost certainly explains the reduction by
interferon-γ (IFN-γ) stimulation of macrophages, shifting polarization azithromycin pretreatment of the serum titre of specific anti-PCV7,
of activated macrophages towards the alternative/anti-inflammatory IgG1 antibodies in mice immunized with polyvalent, polysaccharide,
M2-phenotype, which plays a role in directing Th-2 responses and coor- pneumococcal conjugate vaccine (PCV7) (Fernandez et al., 2004). This
dination of repair following inflammation (Murphy et al., 2008a; finding also indicates that treatment of individuals with azithromycin
Yamauchi et al., 2009). Enhancement of M2-like macrophage genera- should be avoided when an immunization procedure is planned.
tion is observed in vivo after azithromycin pretreatment of mice with In human natural killer (NK) cells stimulated with IL-15, azithro-
a cystic fibrosis-like disease (Legssyer et al., 2006). mycin inhibited cytotoxic function through down-regulation of perforin
The modulation by azithromycin of monocytes/macrophages has expression (Lin et al., 2012). Inflammatory cytokine production was not
been studied in detail in classically activated human blood monocytes inhibited, but it was in the transformed NK cell line, NK-92.
primed with IFN-γ and activated with LPS (Vrancic et al., 2012). It should be noted that the pleiotropic effects of azithromycin are not
Azithromycin, in a concentration-dependent manner, distinctively limited to a single anatomic site or organ or by the presence of microbes.
inhibited the gene expression and/or release of M1 macrophage For instance, azithromycin, like dexamethasone, inhibits the production
markers (CCR7, CXCL11 and IL-12p70), but not TNFα or IL-6 and en- of the acute phase protein serum amyloid-A (SAA) during sterile inflam-
hanced expression and/or release of M2 macrophage markers (IL-10 mation in mice (Glojnaric et al., 2007). SAA, like C-reactive protein
and CCL18), and the pan-monocyte marker CD163. Modulation of the (CRP), is a pentraxin synthesized by hepatocytes after IL-1β or IL-6
Toll-like receptor (TLR) 4 signalling pathway, NF-kB and signal trans- stimulation, and can be inhibited indirectly by azithromycin through re-
ducer and activator of transcription-1 (STAT-1) activation were shown duction of cytokine release at the inflamed site. Nevertheless, given the
to be involved. The inhibitory profile of azithromycin was similar to strong accumulation of azithromycin in hepatocyte lysosomes, it may
that of the lysosomotropic drug chloroquine, indicating that lysosomal also directly inhibit hepatic pentraxin and cyto-/chemokine synthesis
accumulation of the macrolide played a crucial role (see Section 4.2.1). (Matijasic et al., 2012).
In dendritic cells, azithromycin modulates differentiation and LPS-
induced maturation towards a regulatory phenotype (e.g. enhanced 4.2. Cell signaling processes
IL-10) with increased phagocytic capacity (Polancec et al., 2012;
Sugiyama et al., 2007). Furthermore, azithromycin inhibits expression The mechanisms of the immunomodulatory actions of macrolide
of co-stimulatory (CD40 and CD86) and major histocompatibility antibiotics, including azithromycin have been reviewed previously

Table 3
Recent studies on the anti-inflammatory/immunomodulatory effects of azithromycin in animal models in vivo.

Reference Model Azithromycin dosing Variables changed

(Tsai et al., 2004) Murine P.aeruginosa lung infection s.c. from day 3 ↓ lung leukocytes, inflammatory cytokines,
not bacterial counts
(Legssyer et al., 2006) CFTR-mutated mouse P.aeruginosa p.o. 4 wks ↓ lung leukocytes, MPO,
lung infection Pretreatment ↑ lung M2-like macrophages
(Tsai et al., 2009) Murine P.aeruginosa lung infection s.c. from days 1 to 5 ↓ lung bacterial counts, inflammatory cytokines,
↑ survival, lung efferocytotic macrophages
(Yamada et al., 2013) Murine MDR A.baumanii pneumonia s.c. from days 1 to 6 ↓ mortality, lung leukocytes, inflammatory
cytokines, not bacterial counts
(Beigelman et al., 2010) Murine Sendai virus pneumonia s.c. from days 1 to 7 ↓ lung leukocytes, inflammatory cytokines, not viral load
(Tong et al., 2011) Murine Bacillus pyocyaneus shock 1 h before ↑ survival
(Geudens et al., 2008) Murine lung ischemia–reperfusion p.o. 14 days ↓ lung leukocytes, inflammatory cytokines,
injury Pretreatment
(Ballard et al., 2007) Neonatal mouse lung hypoxia From days 1 to 10 ↓ mortality, lung leukocytes, inflammatory
cytokines,
(Terao et al., 2003) Murine i.p. LPS inflammation i.g. 20 min ↓ lung NO, plasma NO & inflammatory cytokines,
(Hao et al., 2013) Pretreatment ↓ plasma kynurenine & inflammatory cytokines,
(Ivetic Tkalcevic et al., 2006) Murine i.p. or i.v. LPS shock p.o. 30 min before ↓ plasma TNFα, mortality
(Bosnar et al., 2009; Bosnar et al., 2011) Murine i.n. LPS lung inflammation p.o./i.p.4 h ↓ lung leukocytes, inflammatory cytokines
Pretreatment IL-1β from alveolar macrophages
(Fernandez-Robredo et al., 2013) Rat subconjunctival LPS 3 days topical pretreatment ↓ ocular inflammation
(Glojnaric et al., 2007) Murine sterile s.c. AgNO3 inflammation p.o. from 1 to 3 days ↓ plasma SAA & inflammatory cytokines,
(Beigelman et al., 2009) Murine OVA asthma s.c. from 1 day before or after challenge ↓ lung inflammation, BAL leucocytes & Th2 cytokines
(Ivetic Tkalcevic et al., 2012) Murine acute OXA skin DTH Topical after challenge ↓ ear edema, leukocytes, IL-4
Murine chronic skin DTH, repeated OXA Topical after each challenge ↓ ear mast cells, IgE, IL-4, not leukocytes,
(Plesko et al., 2010) Murine DNFB colon DTH From before or after challenge ↓ colitis
(Iwamoto et al., 2013) Murine allogeneic GvH disease p.o. from day −2 to 2 ↑ graft resistance, due to effect on DCs?
(Fernandez et al., 2004) Murine immunization to pneumococcal i.p. 1 h before ↓ specific anti-PCV7 IgG1 antibodies
vaccine PCV7

CFTR: cystic fibrosis transmembrane conduction regulatory protein; DNFB: dinitrofluorobenzene; GvH: graft versus host disease to transplanted bone marrow; i.g.: intragastric; i.n.:
intranasal; i.p.: intraperitoneal; i.v.: intravenous; MDR: multidrug resistant; MPO: myeloperoxidase; OVA: ovalbumin; OXA: oxazolone; p.o.: oral; s.c.: subcutaneous.

Please cite this article as: Parnham, M.J., et al., Azithromycin: Mechanisms of action and their relevance for clinical applications, Pharmacology &
Therapeutics (2014), http://dx.doi.org/10.1016/j.pharmthera.2014.03.003
M.J. Parnham et al. / Pharmacology & Therapeutics xxx (2014) xxx–xxx 7

(Kanoh & Rubin, 2010; Shinkai et al., 2008). Mainly based on studies on action of azithromycin on fibroblast MPR300 (and/or on membrane
mucus and cytokine production by airway epithelial and inflammatory phospholipid fluidity) may account for its inhibition of the pinocytosis
cells, the authors concluded that macrolide antibiotics, including of macromolecules and their transport from the plasma membrane to
azithromycin, exert biphasic stimulatory then inhibitory effects on endo/lysosomes (Tyteca et al., 2001). Acute effects on MPR300 may
MAPK kinase (ERK, JNK, p38) activity, with subsequent modulatory ef- also explain the paradoxical increase by azithromycin of lysosomal hy-
fects on NFκB and AP-1 transcription factor activities. Crosstalk between drolase activity in fibroblasts (Gerbaux et al., 1996) and the early release
ERK and p38 was suggested to cause mutually counteracting effects. and prolonged depletion of circulating neutrophil lysosomal enzymes
Azithromycin and to a less extent other macrolide antibiotics, interacts after azithromycin treatment of human volunteers (Culic et al., 2002).
with phospholipids with pronounced effects on lysosomal function and In this context, inhibition of neutrophil lysosomal enzyme release by
these must be incorporated into any consideration of the intracellular neutrophils was the first anti-inflammatory effect of azithromycin
effects of azithromycin. ever to be described (Carevic & Djokic, 1988). Interestingly, Niemann–
Pick C2 protein (NPC2), which transports cholesterol, glycolipids and
4.2.1. Interactions with phospholipids and lysosomes lysophosphatidic acid out of lysosomes, is also transported to endo/
Azithromycin and other macrolides readily penetrate into phospho- lysosomes from the Golgi apparatus by MPR300 (Willenborg et al.,
lipid bilayers, the N′-3 of the desosamine being inserted into the hydro- 2005), so azithromycin could potentially affect extra-lysosomal lipid
phobic domain and the macrocycle occupying a position at the transport. In fact, expression of both NPC1 and NPC2 and of various
hydrophobic–hydrophilic interface (Montenez et al., 1996, 1999). lipid and cholesterol metabolism genes in human epithelial cells was
For azithromycin, this position allows the positively-charged amino increased by incubation with azithromycin (Ribeiro et al., 2009). Since
group in the macrocycle to neutralize the negatively charged groups lysophosphingolipids, like azithromycin (see Sections 4.1.2 and 4.3.5),
in the phosphate head of the acidic phospholipids, including phos- modify macrophage differentiation (Yamamoto et al., 2011), they
phatidylinositol (PI), preferentially found at the inner surface of the could mediate the actions of azithromycin.
cell membrane bilayer (Montenez et al., 1999). These negatively
charged phospholipids bind signaling proteins to the inner plasma-
lemma and regulate many cellular functions, including phagocytosis 4.2.2. Autophagy
(Yeung & Grinstein, 2007). Consequently, neutralization by azithro- Autophagy is a process whereby ubiquitin-tagged or Hsp-
mycin of the negative charge could conceivably initiate many cellular chaperoned, often damaged, cytosolic components or intracellular
effects. pathogens are engulfed in autophagosomes and transported to
PI contains high amounts of arachidonic acid, and in LPS-stimulated lysosomes for degradation (Randow & Munz, 2012). Fusion of
macrophages, azithromycin, like an inhibitor of cytoplasmic phospholi- autophagosomes with lysosomes is mediated by the delipidated
pase A2 (cPLA2), inhibited arachidonate release and eicosanoid produc- autophagy protein, microtubule-associated protein 1A/1B-light
tion but not cPLA2 expression, probably preventing cPLA2 access to its chain 3 (LC3), which binds to phospholipids from apopototic parti-
membrane substrates (Banjanac et al., 2012). Reported inhibition by cles phagocytosed by leukocytes, thereby facilitating the engulf-
macrolide antibiotics, including azithromycin, of arachidonate release ment of the resulting efferocytotic vesicles by lysosomes (Florey
and metabolism in LPS-stimulated neutrophils and mononuclear cells & Overholtzer, 2012). In smooth muscle cells (SMC), azithromycin,
(Miyazaki et al., 2003; Sato et al., 2007) was probably due to effects on at a clinically relevant concentration (10 μM), induced autophagy
signaling further downstream from TLR4. and inhibited actively proliferating cells (Stamatiou et al., 2009).
Interaction of azithromycin with phospholipid head groups This autophagy-mediated inhibition of SMC proliferation may
decreases their mobility and the fluidity and elasticity of the plasma contribute to respiratory effects of azithromycin.
membrane (Fa et al., 2007; Tyteca et al., 2003), accounting for many In contrast, in fibroblasts and human macrophages, azithromycin
effects of the macrolide on cellular functions. These include altered blocks autophagy, measured by accumulation of lipidated LC3 (LC3-
membrane insertion of markers into macrophage membranes and II), by reducing acidification of autophagosomes and lysosomes
slowing of their trafficking into lysosomes; inhibition of fluid phase en- (Renna et al., 2011). Particle-induced generation of TNFα and intra-
docytosis, but not phagocytosis, of macro-molecules; down-regulation cellular mycobacterial killing by the macrophages was inhibited. In
and delayed recycling of surface transferrin receptors, but not Fcγ re- cystic fibrosis (CF) patients treated for 6 months with azithromycin,
ceptors (Tyteca et al., 2002, 2003). Moreover, anti-inflammatory effects mycobacterial killing was also reduced and LC3-II increased (Renna
of a range of synthetic macrolide derivatives on macrophages were et al., 2011). The clinical significance of this finding is questionable
correlated with their phospholipid binding properties and number of because, in a nested case–control study on chronic azithromycin
positively charged centers (Munic et al., 2011). treatment in a total of 27,112 CF patients, the longer the treatment
Extensive lysosomal accumulation of azithromycin results from with azithromycin, the less likely were patients to have non-
proton-driven concentration of its dicationic form in the acidic (pH 5) tuberculous mycobacterial infections (Binder et al., 2013).
lysosomal milieu. Here, azithromycin, like other cationic amphiphilic Azithromycin effects on autophagy may depend on the intracel-
drugs (CAD), inhibits lysosomal phospholipase A1 by blocking phos- lular site of action. Accumulation of azithromycin in muscle cells is
pholipid substrate access, causing phospholipidosis, the accumulation many-fold less than in fibroblasts and leukocytes, in which accumu-
of phospholipids in lysosomes (Kosol et al., 2012; Van Bambeke et al., lation is predominantly lysosomal (Matzneller et al., 2013). Effects
1996). Phospholipidosis, in turn, correlates with phospholipid binding on SMC, thus, may involve surface membrane effects, while in the
by a range of macrolide derivatives (Munic et al., 2011). other two cells, azithromycin may target predominantly lysosomal
Annother mechanism of action may also account for phospho- degradation processes. Macrolides bind with low affinity to
lipidosis. Following phosphorylation of their mannose residues in the valosin-containing protein (VCP), a crucial mediator of intracellular
Golgi apparatus, lysosomal enzymes are transported to endosomes by protein degradation and autophagy (Ju & Weihl, 2010; Nujic et al.,
the multifunctional mannose-6-phosphate/insulin growth receptor II 2012b; Tresse et al., 2010) which may account for some long-term
receptor, MPR300 (Kornfeld, 1992). Here, the acidic endosome environ- effects of azithromycin on autophagy.
ment causes enzyme release and transfer to lysosomes, while MPR300 Recently, azithromycin, clarithromycin and erythromycin
is returned to the Golgi apparatus. All phospholipidosis-inducing CAD were found to sensitize to cytotoxic and pro-apoptotic effects
tested, including erythromycin, prevent MPR300 recycling from the of bortezomib by blocking autophagic flux in myeloma cells
endosomes to the Golgi apparatus, causing lysosomal depletion and ex- (Moriya et al., 2013), suggesting a potential use for macrolides
tracellular secretion of lysosomal enzymes (Ikeda et al., 2008). Such an in oncology.

Please cite this article as: Parnham, M.J., et al., Azithromycin: Mechanisms of action and their relevance for clinical applications, Pharmacology &
Therapeutics (2014), http://dx.doi.org/10.1016/j.pharmthera.2014.03.003
8 M.J. Parnham et al. / Pharmacology & Therapeutics xxx (2014) xxx–xxx

4.2.3. Apoptosis 4.2.5. NFκB pathway


Apoptosis is a crucial non-inflammatory process for terminating the In no study on MUC5AC secretion by airway epithelial cells was an in-
life of cells, particularly immunocompetent cells. Interactions exist be- hibitory effect of azithromycin observed on NFκB activation, despite
tween apoptosis and autophagy in that the proteins, autophagy protein clear NFκB activation by all stimuli (Araki et al., 2010; Morinaga et al.,
5 (Atg5) and autophagy inducing protein, Beclin-1 are involved in both 2009; Nie et al., 2012). However, apart from CF cells, a clear correlation
processes (Fimia & Piacentini, 2010). It is, therefore, not surprising that exists, in epithelial cells, between inhibition of IL-8 generation, NFκB ac-
effects of azithromycin have been observed on apoptosis. tivation and DNA binding (Cigana et al., 2006; Matsumura et al., 2011;
In SMC, azithromycin initially induces autophagy (Stamatiou et al., Saint-Criq et al., 2012). This (indirect) NFκB inhibition involves an
2009) and then apoptosis after prolonged incubation (Stamatiou et al., action on the receptor signal-transmitting Rho GTPase, Ras-related C3
2010). At concentrations of 100 μM or higher, azithromycin induces botulinum toxin substrate 1 (Rac-1) (Matsumura et al., 2011). Thus,
apoptosis in neutrophils (Koch et al., 2000), an action shared by many while azithromycin acts on epithelial mucin by targeting AP-1, inhibi-
antibiotics (Healy et al., 2002); in human lymphocytes by inhibiting ex- tion of IL-8 involves both a direct action on AP-1 and indirect on NFκB
pression of Bcl-xL (Ishimatsu et al., 2004; Mizunoe et al., 2004); and in activation, apparently preceded by inhibition of membrane Rac-1,
NK cells (Lin et al., 2012). Induction of apoptosis is thus related to supporting an action of azithromycin close to the phospholipid cell
prolonged, extensive accumulation, particularly at high doses, as seen membrane.
with circulating human neutrophils (Culic et al., 2002). This likelihood It is unclear whether NFκB is a potential target for azithromycin in-
is supported by studies on monocytes and macrophages. hibition of leukocyte cytokine production (Aghai et al., 2007; Ikegaya
Despite accumulating several hundred-fold in macrophages, et al., 2009). Several macrolide antibiotics, though, inhibit NFκB activa-
azithromycin does not readily induce apoptosis in these cells. To the tion in a variety of cells in vitro (Kanoh & Rubin, 2010), particularly
contrary, in human bronchial alveolar lavage (BAL) macrophages, lyso- when the incubation period is long (Shinkai et al., 2008) and upstream
somal membrane permeabilization and ROS-induced cell death were signaling has occurred. Moreover, in classically activated human mono-
inhibited by long-term azithromycin treatment, an action confirmed cytes, the macrolide only reduced NFκB activation at relatively high
in vitro (Persson et al., 2012). Macrophages from azithromycin-treated concentrations (50 μM) (Vrancic et al., 2012). Since both AP-1 and
patients, in fact, show enhanced ability to phagocytose apoptotic cells. NFκB are both dependent on upstream signaling processes, many of
First observed in patients with chronic obstructive pulmonary disease the observations on their inhibition by azithromycin may be secondary
(COPD), azithromycin in vitro restored to normal the reduced ability to direct cell membrane effects of the macrolide (see Section 4.2.1).
of alveolar macrophages from COPD patients to phagocytose apoptotic
neutrophils (Hodge et al., 2006). Administered for 12 weeks to COPD 4.2.6. Gene expression
patients, the macrolide also enhanced alveolar macrophage effero- An overview of changes induced by azithromycin, at clinically rele-
cytotic capacity and mannose receptor (CD 206) expression, together vant concentrations, was gained from human airway epithelial cells
with a decrease in epithelial cell apoptosis (Hodge & Reynolds, 2012; stimulated with CF lung supernatant (Ribeiro et al., 2009). In confirma-
Hodge et al., 2008). This enhancement of macrophage efferocytosis is tion of other studies, genes for mucin production, MMPs, CXCL3, CXCL6,
now attributed to the ability of azithromycin to promote the expression IL-7, tissue plasminogen activator (tPA) and cell cycle were down-
of the M2 macrophage phenotype (Section 4.1.2). regulated. Genes for CXCR7, IL-1ra, IL-6 signal transducer, transferrin
receptor (CD71) and VEGFα were upregulated. IL-8 expression and se-
4.2.4. MAP kinases and AP-1 cretion was increased only by short-term (6 h) incubation. In keeping
Kanoh and Rubin (2010) stated in their review that“there have been with effects of azithromycin on lipid metabolism, the expression of
so many different reported effects of macrolides on cell signaling path- many proteins regulating lipid and cholesterol metabolism were up-
ways that attempts to put these actions together as one mechanism is regulated. Prominent among these were SREBP1 (sterol regulatory ele-
difficult. In fact, all macrolides seem not to have similar ranges or de- ment binding protein 1) and SREBP2, master activating transcription
grees of activity.” In a majority of the reports, clarithromycin was the factors which control the expression of many other genes for cholesterol
macrolide tested, so it is important to consider azithromycin separately. metabolism. Effects of azithromycin on membrane phospholipids
The clearest indication for involvement of the Erk1/2 MAP kinase and SREBP1/2 may be linked, SREBP1 phosphorylation and down-
signalling pathway has been obtained from studies on mucin produc- regulation of cell cycle genes being regulated by effects on MAP kinases.
tion by airway epithelial cells. In these cells, azithromycin inhibited In summary, direct effects of azithromycin on charged phospholipds
mucin 5 AC (MUC5AC) secretion and the increase in Erk1/2 phosphory- in plasma and lysosomal membranes seem to be initial targets of the an-
lation induced by some, but not all stimuli (Ishimotoet al., 2009; Luo tibiotic, with associated changes in the expression of lipid metabolising
et al., 2011; Morinaga et al., 2009). Subsequent transcription of the genes, such as SREBP, autophagy and apoptosis. Effects on Rac-1, MAPK
MUC5AC gene, in H. influenzae-or EGF-stimulated cells, was inhibited kinases and other enzymes activated by membrane changes probably
by azithromycin through inhibition of AP-1 activation (Araki et al., determine downstream effects on transcription factors AP-1 and NFκB,
2010; Nie et al., 2012). The same signaling pathway is linked with inhib- depending on the cell type and membrane stimulus (Fig. 1). Additional
itory effects of azithromycin on IL-8 production in epithelial cells direct effects of azithromycin on signaling molecules cannot be ruled
(Srivastava et al., 2011) and on human neutrophil chemotaxis (Tsai out. However, only valosin containing protein (VCP), a protein involved
et al., 2004). in autophagy, has been identified as a low affinity (KD = 1.8 × 10−4 M)
Subsequent inhibition of AP-1 binding to DNA by azithromycin has azithromycin binding protein (Nujic et al., 2012b). No high affinity bind-
been associated with inhibition of IL-8 production in CF airway cells ing proteins for azithromycin were detected.
(Cigana et al., 2006); inhibition of proliferation in human vascular
SMCs, (Miller et al., 2000); suppression of IL-12p40 expression and 4.3. Role of state and phase of inflammation
generation by M1-type macrophages (Yamauchi et al., 2009); and
inhibition of IL-1β production in LPS-stimulated macrophages (Bosnar Results of several investigations indicate that the immunomodulato-
et al., 2011). Clearly, for inhibition of mucin and cytokine production ry effects of azithromycin and other macrolide antibiotics vary accord-
and of neutrophil chemotaxis, the ERK1/2, AP-1 signaling pathway is ing to the presence and phase of an inflammatory response.
an important target of azithromycin. The effect on AP-1 binding, though,
is probably indirect as basal, unstimulated AP-1 binding and IL-8 4.3.1. Distinction between non-inflamed and inflammatory conditions
production is not affected by azithromycin (Cigana et al., 2006; Miller The activation state of immune cells determines their response
et al., 2000). to azithromycin. Thus, the macrolide antibiotics, as a class, stimulate

Please cite this article as: Parnham, M.J., et al., Azithromycin: Mechanisms of action and their relevance for clinical applications, Pharmacology &
Therapeutics (2014), http://dx.doi.org/10.1016/j.pharmthera.2014.03.003
M.J. Parnham et al. / Pharmacology & Therapeutics xxx (2014) xxx–xxx 9

neutrophil and macrophage responses in healthy rodents, but generally mononuclear phagocyte, promoting an inflammation-resolving M2
inhibit inflammation and leukocyte responses in experimental inflam- macrophage phenotype or a regulatory DC phenotype.
matory models (Culic et al., 2001). This dichotomy was clearly seen in
a study in which pretreatment with azithromycin had no effect on 5. Clinical applications in
lung inflammatory cytokines, but increased IL-10 concentrations in respiratory infections and airway inflammation
healthy mice, while in mutant Δ508 CF mice, azithromycin inhibited
lung inflammation and pro-inflammatory cytokines, but had no effect The unique imunomodulatory properties of azithromycin, its broad
on IL-10 (Legssyer et al., 2006). This may have been due to effects on antibacterial spectrum and exceptional pharmacokinetics, resulting
predominantly resident macrophages in healthy and M1-like macro- in extensive and sustained tissue penetration, largely account for its
phages in inflamed CF mice. disease-modifying effects in infectious or non-infectious inflammatory
This difference between healthy and inflamed conditions is also seen airways diseases. These coincident properties also explain its indication
in humans. Standard dosing of healthy volunteers with azithromycin for for upper and lower respiratory tract infections (e.g. acute bacterial
3 days stimulated blood neutrophil degranulation and oxidative burst, sinusitis, community-acquired pneumonia), in which macrolides exert
but not in similarly treated COPD patients (Culic et al., 2002; Parnham therapeutic benefits not solely explainable by antibacterial activities
et al., 2005). In untreated endothelial cells, azithromycin increases (Amsden, 2005).
(Millrose et al., 2009), whereas in LPS-stimulated epithelial cells or neu-
trophils, macrolides suppressed adhesion molecule expression (Khair 5.1. Upper and lower respiratory tract infections
et al., 1995; Lin et al., 2000). It seems likely that administered early dur-
ing a bacterial infection, azithromycin could promote host defense by With a broad antibacterial spectrum and oral bioavailability,
activating naïve leukocytes and endothelial cells, unprimed by cyto- macrolides are the second most commonly used antibiotic class in
kines. At a later stage, when cells are activated by inflammatory stimuli, many European countries (Adriaenssens et al., 2011), exceeded only
immunomodulation by azithromycin would lead to inhibition of inflam- by penicillins. Based on their efficacy against the main respiratory
mation and promotion of its resolution (Parnham, 2005). pathogens (S. pneumoniae, S. pyogenes), macrolides are often used for
empirical treatment of respiratory tract infections. They are also active
against other bacteria including Streptococcus pneumoniae, Mycoplasma
4.3.2. Time dependency of effects
pneumoniae, Chlamydia spp., Legionella pneumophila, Niesseria
The concept of a biphasic, time-dependent immunomodulatory
gonorrhoeae, Bordatella pertussis, Campylobacter jejuni, Treponema
action of azithromycin and other macrolide antibiotics is supported by
pallidum, Ureaplasma urealyticum, Corynebacterium diphtheria and
data in several publications. Initial stimulatory and later enhancing
Listeria monocytogenes. Azithromycin has an even broader spectrum,
effects have been observed with clarithromycin on IL-8 production
being active against Haemophilus influenza, Moraxella catarrhalis, non-
and Erk1/2 activation in bronchial epithelial cells (Shinkai et al., 2006)
tuberculosis mycobacteria, Bartonella henselae, Rhodococcus equi,
and with erythromycin in vitro or roxithromycin given to mice on
Toxoplasma spp., Cryptosporidium spp. and Plasmodium species
mononuclear cell IL-1 and IL-2 production (Kita et al., 1990; Konno
(Blondeau et al., 2002). H. influenzae, M. catarrhalis, S. pneumoniae,
et al., 1992). Similarly, in Candida albicans infected mice and in healthy
as well as Chlamydophila pneumonia and M. pneumonia are particularly
human volunteers, a merely transient stimulation of neutrophil chemo-
susceptible. Studies in mice indicate that even with multidrug re-
taxis was seen with erythromycin (Fernandes et al., 1984), a finding also
sistant bacterial strains, clinical benefit may result from the immuno-
obtained with azithromycin in healthy human volunteers (Culic et al.,
modulatory actions of the drug (Yamada et al., 2013).
2002). In the latter study, initially enhanced blood neutrophil activation,
Azithromycin does not inhibit growth of Pseudomonas aeruginosa
2.5 h after azithromycin, changed to a significant reduction after 1–
in vitro, but shows efficacy in vivo mainly because of inhibition of
28 days.
quorum-sensing and biofilm (Hoiby, 2011) (see Section 2). Given
Early, mainly stimulatory effects of azithromycin, especially on
prophylactically, azithromycin, therefore, reduces the incidence
neutrophils, are probably mediated by interactions with phospholipids
of ventilation-associated pneumonia, particularly in patients with
and Erk1/2, followed by later modulation of transcription factors AP-1
rhamnolipid-positive isolates (van Delden et al., 2012), confirming
and NFκB. Subsequent delayed inhibitory effects on cell function are
previous findings in P. aeruginosa-induced lung inflammation in mice
likely related to lysosomal accumulation of the macrolide, with disrup-
(Table 3). Promotion by azithromycin of macrophages with a regulatory
tion of protein and lipid transport through the Golgi apparatus and ulti-
M2-like phenotype in animals (Feola et al., 2010; Hoffmann et al., 2007)
mate effects on surface receptor expression, including macrophage
suggests that immunomodulation also contributes to the benefit seen
phenotype changes and autophagy (see Section 4.2) (Fig. 2).
clinically in P. aeruginosa infection.
Clinical experience with azithromycin in the treatment of upper
4.3.3. Resolution of inflammation and lower respiratory tract infections has been reviewed previously
Since azithromycin can drive the generation of an anti-inflammatory (Mulholland et al., 2012; Panpanich et al., 2008; Wierzbowski et al.,
M2 macrophage phenotype (Section 4.1.2), it is pertinent that it also 2006). Accumulation in leukocytes at inflamed sites means that
promotes the resolution of experimental inflammation. In Δ508 CF oral dosing either at 500 mg daily for 3 days or as a single dose of 1–
mice with LPS lung inflammation, azithromycin pretreatment inhibited 1.5 mg is often sufficient.
inflammation 48 h after LPS, when lung IL-10 levels were increasing, i.e. Azithromycin is effective in microbial eradication in patients
during the resolution phase (Legssyer et al., 2006). This effect on the with presumed bacterial lower respiratory tract infections (LRTI),
resolution process was further clarified in murine, self-limiting, community-acquired pneumonia and acute bacterial sinusitis (Henry
zymosan-induced, peritoneal inflammatory cell infiltration, which et al., 2003; Lakos et al., 2012; Panpanich et al., 2008; Yanagihara
peaks at 48 h and is inhibited by azithromycin pretreatment (Navarro- et al., 2009). In acute sinusitis, there are few randomized studies of
Xavier,et al., 2010). Administered at 72 h, azithromycin inhibited azithromycin, but the clinical and bacteriologic efficacy and safety of
remaining leukocyte infiltration and also enhanced the ratio of M2/M1 azithromycin is generally good and comparable to or even slightly
macrophages in the peritoneal exudate. In contrast to other anti- better than that of amoxicillin/clavulanate (Marple et al., 2010) or
inflammatory drugs, azithromycin did not decrease the modulatory levofloxacin (Murray et al., 2005). Azithromycin therapy was better
cytokine, IL-10, in exudates. These data strengthen the proposal that tolerated, with less frequent adverse reactions and lower cost, com-
apart from acute anti-inflammatory effects on neutrophils, the major pared to amoxicillin/clavulanate (Henry et al., 2003; Karpov, 1999). A
target for the immunomodulatory effects of azithromycin is the common clinical finding is the coexistence of sinusitis, periodontal

Please cite this article as: Parnham, M.J., et al., Azithromycin: Mechanisms of action and their relevance for clinical applications, Pharmacology &
Therapeutics (2014), http://dx.doi.org/10.1016/j.pharmthera.2014.03.003
10 M.J. Parnham et al. / Pharmacology & Therapeutics xxx (2014) xxx–xxx

Phospholipid
bilayer penetration Lysosomal
enzyme &
Inflammatory
ROS release
stimulus
Neutralizes negative
PLs, reduces FA release,
decreases membrane fluidity

Lipid Rac-1?
Surface remodeling
molecule MAPKs (ERKs)
recycling
Lysosomal
accumulation Phospholipidosis NFκB SREBP1
AP-1
Increased Lysosome
phagocytosis? +LC3-II
High concs
+VCP
+ MPR
Autophagy - Bcl
Enzymes
Apoptosis

Lipid Golgi
transport
Cytokine, mucin and
ER inflammatory gene expression
depending on stimulus

Fig. 1. Effects of azithromycin on cell signaling. Azithromycin penetrates the cell membrane bilayer and stabilizes the membrane, reducing fluidity. The cationic moiety of azithromycin
comes into close proximity to the negatively charged phospholipids in the inner leaflet of the membrane, neutralising their charge. This results in reduced fatty acid release and also to
liberation of enzymes bound by electrostatic charge to the membrane. Lipid remodeling leads to modulation of signaling pathways, particularly of MAPK kinases, such as ERK1/2 (possibly
involving Rac-1) with subsequent inhibition of the activation of transcription factors including AP-1 and NFκB. The signaling pathway most affected is likely to be dependent on the
particular cell, its activation state and the stimulus by which it is activated. Lysosomal enzyme release and initial induction of the oxidative burst of neutrophils are a consequence of
the actions on membrane lipids. Membrane lipid remodelling also disrupts surface molecule recycling and probably phagocytosis. Molecules dependent on negatively charged
phospholipids are also affected, including LC3-II, with subsequent influences on lysosomal uptake and autophagy. The latter may also be influenced, with time, by low affinity binding
of azithromycin to VCP. Azithromycin accumulates in lysosomes, modulating MPR transport of enzymes and lipids and through lipid remodeling in the lysosome membranes ultimately
leads to phospholipidosis. Apoptosis may be induced by high concentrations of azithromycin, possibly by compromising the action of the apoptosis inhibiting Bcl molecules. [Ø indicates
inhibition].

disease and/or bronchiectasis (Brook, 2006), which may all benefit (Branden et al., 2004; Esposito et al., 2005; Mulholland et al., 2012).
simultaneously from azithromycin therapy (Sections 5.3 and 8.2). Long-term, low-dose azithromycin, though, lacked efficacy in children
Immunomodulation by macrolides has been proposed repeatedly to with acute respiratory syncytial virus bronchiolitis (Kneyber et al.,
contribute to decreased length of hospitalization and increased survival 2008; Pinto et al., 2012) or in chronic rhinosinusitis (Videler et al.,
in CAP and CAP-related sepsis (Amsden, 2005; Giamarellos-Bourboulis 2011). Nevertheless, an increasing number of small scale studies sug-
et al., 2008; Metersky et al., 2007; Restrepo et al., 2009; Tessmer et al., gest that azithromycin and other macrolides are of benefit in viral infec-
2009). Recent meta-analyses indicate that while macrolides do decrease tions because of their immunomodulatory properties (Min & Jang,
mortality among hospitalized CAP patients, in RCTs or in guideline- 2012). There is considerable interest, particularly in Japan, in using
concordant studies a difference from other antibiotics was not evident macrolides to prevent the cytokine storm which occurs in patients dur-
(Asadi et al., 2012; Asadi et al., 2013; Skalsky et al., 2013). As discussed ing respiratory viral epidemics (Bermejo-Martin et al., 2009). Certainly,
later for sepsis (Section 7), disease severity may determine sensitivity to azithromycin inhibits and stimulates the resolution of lung inflamma-
antibiotics. A recent study on 187 patients with pneumonia due to tion in mice infected with mouse parainfluenza virus type 1 (Sendai
S. pneumoniae, clearly confirmed that azithromycin significantly virus) without affecting viral load (Beigelman et al., 2010). It may well
reduces mortality rate (Shorr et al., 2013). This study differed from also enhance anti-viral defences, since in contrast to erythromycin and
previous studies in being limited to a single causative agent. telithromycin, azithromycin significantly increases interferon genera-
Azithromycin effectively reduces respiratory symptoms in chronic tion and interferon-stimulated gene expression, reducing virus replica-
or recurrent respiratory tract infections due to so-called atypical bacte- tion and growth in human bronchial epithelial cells infected with
ria, such as chronic Chlamydia pneumoniae or Mycoplasma pneumoniae rhinoviruses 16 and 1B (Gielen et al., 2010).

Please cite this article as: Parnham, M.J., et al., Azithromycin: Mechanisms of action and their relevance for clinical applications, Pharmacology &
Therapeutics (2014), http://dx.doi.org/10.1016/j.pharmthera.2014.03.003
M.J. Parnham et al. / Pharmacology & Therapeutics xxx (2014) xxx–xxx 11

5.2. Diffuse panbronchiolitis 5.3. Cystic fibrosis and non-cystic fibrosis bronchiectasis

Initial reports on clinical immunomodulation by macrolides ap- As in DPB, pulmonary neutrophil infiltration predominates in
peared in the late 1980s and 1990s, when it became clear that the patients with CF and non-CF bronchiectasis and this encouraged
prognosis of patients with diffuse panbronchiolitis (DPB) improved dra- many investigators to study azithromycin in this indication. In an RCT
matically after implementing long-term treatment with macrolides. in 251 patients chronically infected with P. aeruginosa, standard 3 day
This was first documented with erythromycin, demonstrating signifi- azithromycin treatment weekly for 24 weeks significantly improved
cant reduction in broncho-alveolar lavage (BAL) fluid of IL-1β, IL-8 FEV1 and reduced risk of exacerbations (Saiman et al., 2003). Sub-
and neutrophils (Sakito et al., 1996). Sputum volume in DPB is reduced sequently, when treatment was extended to 12 months, FEV1 remained
and lung function improved by macrolides, with marked reduction of unchanged in patients not infected with P. aeruginosa, although exacer-
neutrophil activation and lung infiltration, of pro-inflammatory cyto- bation and cough were significantly decreased (Clement et al., 2006;
kines and MMPs. This benefit is seen despite the relatively high inci- Saiman et al., 2010). While pulmonary function, thus, only improves
dence of resistance to macrolides in Japan (Kobayashi et al., 1993; on azithromycin treatment for up to 6 months and not thereafter, re-
Kudoh, 2004). Because of the established treatment of DPB with other duced exacerbations and reduced S. aureus counts, presumably related
macrolides in Japan, azithromycin has been used less frequently, but to immunomodulation, are consistent findings (Southern et al., 2011).
was recently reported to be effective and safe for long-term therapy The FEV1 evolution is explained by a decrease in FEV1 during the initial
(Li et al., 2011). inflammatory phase, which then increases with decreasing inflamma-
Treatment of DPB is the most striking example of the introduction of tion after azithromycin is initiated, FEV1 subsequently reaching a
macrolide-therapy to chronic respiratory diseases. However, the experi- steady-state or plateau phase (depending on % predicted for each
ence in DPB is almost exclusively based on non-randomized, controlled individual). Azithromycin is generally considered an important part
clinical trials (RCTs) or retrospective studies (Kudoh et al., 1998; Li et al., of the treatment armamentarium for CF, particularly in patients chron-
2011; Yang et al., 2013). Drawing on this experience in DPB, long-term, ically infected with P. aeruginosa (Spagnolo et al., 2013; Suresh Babu
low-dose azithromycin therapy was subsequently introduced in CF et al., 2013).
(Saiman et al., 2003), non-CF bronchiectasis (Anwar et al., 2008), The role of azithromycin in non-CF bronchiectasis is now well
COPD (Blasi et al., 2010; Hodge et al., 2008; Parnham et al., 2005), established with three recent RCTs. In the EMBRACE trial, compared to
asthma (Piacentini et al., 2007), infectious bronchiolitis (Naidoo & placebo, azithromycin decreased event-based exacerbation rates in
Bryant, 2009) and BOS following lung or hematopoietic stem cell adult patients with non-CF bronchiectasis, whereas changes in FEV1
transplantation (Gerhardt et al., 2003; Gottlieb et al., 2008; Vos et al., from baseline and in SGRQ total score during the 6-month treatment
2010). Clinical evidence for immunomodulatory effects and applica- period did not differ (Wong et al., 2012). The BAT trial confirmed that
tions of azithromycin in inflammatory airway disorders is briefly daily use of azithromycin for 12 months results in a lower rate of in-
discussed here based on landmark and important clinical trials. fectious exacerbations but azithromycin also increased FEV1 by 1.03%

Fig. 2. Proposed time course of azithromycin actions on infection and chronic inflammation.

Please cite this article as: Parnham, M.J., et al., Azithromycin: Mechanisms of action and their relevance for clinical applications, Pharmacology &
Therapeutics (2014), http://dx.doi.org/10.1016/j.pharmthera.2014.03.003
12 M.J. Parnham et al. / Pharmacology & Therapeutics xxx (2014) xxx–xxx

per 3 months in comparison to a decrease of 0.10% per 3 months in the 5.6. Other respiratory conditions
placebo group (Altenburg et al., 2013). Macrolide resistance rate was
88% in azithromycin-treated individuals and 26% in the placebo group. Macrolides (i.e. clarithromycin) appear to modulate cryptogenic
In children with non-CF bronchiectasis, once-weekly azithromycin for organizing pneumonia, radiation-related bronchiolitis obliterans
up to 24 months also decreased pulmonary exacerbations, but again organizing pneumonia (Mann et al., 2005; Stover & Mangino, 2005)
with increased carriage of azithromycin-resistant bacteria (Valery and pulmonary fibrosis (Wuyts et al., 2010). This is not surprising
et al., 2013). Overall, these RCTs confirm the beneficial effects with given that in these conditions, the distal airways and alveoli initially
azithromycin or erythromycin described in prior, smaller, retrospective may be involved in an inflammatory process.
studies in non-CF bronchiectasis (Anwar et al., 2008; Serisier & Martin, In summary, numerous pulmonary disorders sharing common
2011). Although not addressing the underlying etiology of the disease, innate inflammatory mechanisms can potentially benefit from the
azithromycin may be a therapeutic option in patients with frequent anti-inflammatory and immunomodulatory properties of azithromycin.
exacerbations (Spagnolo et al., 2013; Suresh Babu et al., 2013). Azithromycin may, therefore, prove to be a new milestone in the treat-
ment of chronic pulmonary disorders, always taking into account the
increase in macrolide resistance. In this respect, future development of
5.4. Chronic obstructive pulmonary disease and asthma non-antibacterial, anti-inflammatory macrolide derivatives may be
crucial (Bosnar et al., 2012; Kobayashi et al., 2013; Mencarelli et al.,
Short-term azithromycin has been an alternative to first-line antibi- 2011; Tomaskovic et al., 2013) (Erakovic Haber et al., in press).
otics in the treatment of acute exacerbations of COPD (AECOPD) for sev-
eral years (Butorac-Petanjek et al., 2010; Yamaya et al., 2012). The 6. Infections of the urogenital
benefit achieved is greater than with other therapies such as corticoste- tract and sexually-transmitted infections
roids or roflumilast (Spagnolo et al., 2013). Long-term azithromycin
both decreased frequency of exacerbations and enhanced quality of Over 90% of an oral dose of azithromycin is excreted via the hepato-
life in AECOPD, though with a small incidence of hearing decrements biliary system (Ballow et al., 1998; Foulds et al., 1990; Singlas, 1995).
(Albert et al., 2011). Azithromycin given for several months to patients Since very little drug is found unchanged in the urine, azithromycin is
with COPD also increased efferocytosis in airway macrophages, in- not suitable for diseases involving significant bacteruria. Due to its opti-
dicating an immunomodulatory component to its efficacy (Hodge & mal systemic distribution and tissue penetration, though, azithromycin
Reynolds, 2012). accumulates up to 100-fold in a variety of tissues of urological interest
Azithromycin ameliorates bronchial hyperresponsiveness and air- (Foulds et al., 1990; Foulds et al., 1991).
way neutrophil infiltration in some children with asthma (Piacentini
et al., 2007), but failed to reduce rates of severe exacerbations and 6.1. Urethritis
LRTI in adult patients with severe asthma (Brusselle et al., 2013). Corti-
costeroids have limited efficacy in COPD or neutrophilic asthma and in a Their considerable penetration into human cells makes macrolides
subgroup analysis of patients with non-eosinophilic severe asthma, particularly suited for treatment of infections caused by obligate intra-
azithromycin significantly lowered rates of severe exacerbations and cellular pathogens, like Chlamydia spp. For several years, azithromycin
LRTI requiring treatment with antibiotics (Brusselle et al., 2013). has been the first-line drug for treatment of primary non-gonococcal
Modulation of neutrophilic inflammation probably accounts for these urethritis (www.uroweb.org/gls/pdf/15_Urological_Infections.pdf),
findings in asthma since azithromycin inhibits airway inflammation in mainly caused, in both sexes, by Chlamydia trachomatis, Ureaplasma
a non-infectious, ovalbumin-induced mouse model of allergic asthma urealyticum or Mycoplasma genitalium.
(Beigelman et al., 2009) (see Section 4.3.3). Essential clinical evidence was provided in 1995 for this indication
Despite broad inhibitory activity in experimental studies (see in men (Stamm et al., 1995). A single 1 g oral dose of azithromycin,
Section 4.1), efficacy of azithromycin in clinical airways inflammation for empirical treatment of acute non-gonococcal urethritis syndrome,
is limited. Clinical bronchodilatory effects have not been reported. was as effective as a standard 7-day course of twice-daily 100 mg doxy-
Azithromycin has a place in the treatment of COPD exacerbations, but cycline in achieving clinical cure. Microbiological and overall clinical
other than possibly contributing to corticosteroid-sparing in highly cure rates with azithromycin were 83% and 81%, respectively (Stamm
selected patients, its potential role in the treatment of asthma remains et al., 1995).
unclear (Spagnolo et al., 2013; Suresh Babu et al., 2013). A meta-analysis of RCTs confirmed the therapeutic equivalence of
the two regimens for chlamydial urethritis in men and women (Lau &
Qureshi, 2002). However, patient compliance to doxycycline for 7
5.5. Post-transplant bronchiolitis days is often hampered by gastrointestinal side effects (Thorpe et al.,
1996). Since poor compliance is detrimental to resolution of the infec-
In lung transplant recipients with post-transplant bronchiolitis, tion and may increase resistance, in our opinion, the 100% compliance
azithromycin for 3–6 months reduced local airway inflammation (cyto- with a single-dose azithromycin regimen may minimize such risks.
kines, neutrophils), oxidative stress and matrix remodelling (Verleden Whereas azithromycin and doxycycline are equally active against
et al., 2006, 2011a,b). It attenuates both neutrophilia and plasma levels chlamydial infections, the macrolide achieved higher rates of eradica-
of multiple pro-inflammatory chemokines (Federica et al., 2011). These tion and clinical remission in patients with non-gonococcal urethritis
findings with azithromycin have even led to the definition of novel (NGU) caused by Mycoplasma genitalium (Falk et al., 2003; Mena et al.,
phenotypes of post-transplant chronic lung allograft dysfunction, 2009).
based on factors such as severity and degree of neutrophil infiltration In women, chlamydial cervicitis/urethritis is highly prevalent, and
(Gottlieb et al., 2008; Vanaudenaerde et al., 2008a,b; Vos et al., 2010). may occur concurrently with onset of preterm labor, preterm rupture
Interestingly, prophylactic azithromycin treatment following lung of membranes, spontaneous abortion, fetal death, postpartum endome-
transplantation resulted in significantly higher chronic rejection-free tritis, salpingitis, with risks to the neonate (Pitsouni et al., 2007). A sin-
survival and lower airway neutrophilia and systemic CRP levels in gle 1 g dose of azithromycin eradicates the infection in over 95% of cases
comparison to placebo (Vos et al., 2011). This again confirms that and is equivalent in terms of efficacy and symptom resolution to one-
azithromycin attenuates inflammation resulting from allo- and non- week, once-daily 100 mg doxycycline (Thorpe et al., 1996). A meta-
allo-immunologic events in the lung. Long-term RCTs are needed to analysis (Pitsouni et al., 2007), involving 587 pregnant women, demon-
establish the routine use of azithromycin in this condition. strated high and equivalent efficacy of azithromycin, erythromycin and

Please cite this article as: Parnham, M.J., et al., Azithromycin: Mechanisms of action and their relevance for clinical applications, Pharmacology &
Therapeutics (2014), http://dx.doi.org/10.1016/j.pharmthera.2014.03.003
M.J. Parnham et al. / Pharmacology & Therapeutics xxx (2014) xxx–xxx 13

amoxicillin against chlamydial cervicitis/urethritis. However, azithro- It is estimated that severe sepsis and septic shock, mainly due to CAP,
mycin was associated with fewer gastrointestinal adverse events, annually affect 377 of 100,000 members of the general population;
fewer total adverse events, and better compliance. mortality is 35–50% making sepsis a leading cause of death, more com-
mon than malignancies and heart disorders (Kotsaki & Giamarellos-
6.2. Prostatitis Bourboulis, 2012). Current guidelines recommend intravenous antibiot-
ic treatment as early as possible, because with most antibiotic classes,
Chronic bacterial prostatitis (CBP) is a difficult-to-eradicate biofilm effectiveness decreases with time and sepsis severity (Dellinger et al.,
disease, dramatically affecting quality of life and requiring long-term 2013). Several retrospective analyses of prospectively collected data
antibacterial therapy for weeks or even months. Eradication of causative have assessed the outcome of monotherapy over combination therapy
pathogens is hampered by poor penetration of most antibacterial agents of severe CAP. The primary endpoint of these analyses was whether
into the prostate gland. Fluoroquinolones are currently the agents of inclusion of a macrolide (azithromycin, clarithromycin or erythromycin,
choice, due to adequate prostatic penetration, but recent RCTs show where recorded) in the combination affects mortality. The summary of
that macrolides are becoming a primary therapeutic option for CBP the seven studies published after 2006 is shown in Table 4. Four studies
(Kolumbic Lakos et al., 2011; Magri et al., 2007, 2010; Skerk et al., indicated a significant decrease in risk of death with the intake of one
2002, 2003, 2004, 2006). The rationale is the excellent penetration macrolide (Martin-Loeches et al., 2010; Metersky et al., 2007; Restrepo
and high concentration of this macrolide (see Sections 3.1 and 3.2) in et al., 2009; Tessmer et al., 2009); two showed a significant decrease in
prostatic tissue and secretions, together with its powerful anti-biofilm duration of hospitalization (Ambroggio et al., 2012; Wilson et al., 2012);
and immunomodulatory activities. Consequently, azithromycin is and only one failed to show any benefit (Karhu et al., 2013).
recommended as a therapeutic option for CBP in specific cases by Two prospective, RCTs have evaluated the impact of macrolide
the European Urological Association (www.uroweb.org/gls/pdf/15_ administration on the outcome of severe infections. In both trials, the
Urological_Infections.pdf). The clinical and preclinical evidence macrolide was clarithromycin, but in view of the striking outcomes,
supporting azithromycin as a valuable treatment option for CBP has they are discussed here as indicators of the potential use of other
been reviewed in detail (Perletti et al., 2011). macrolides, such as azithromycin. The hypothesis was that clarithro-
mycin, administered at a dose of 1 g within 1 h on consecutive days,
6.3. Other sexually transmitted diseases (STDs) should reach maximum serum levels (10 μg/mL) sufficient to modulate
circulating monocyte function. In the first study (Giamarellos-
Macrolides and azithromycin in particular, are used to treat a num- Bourboulis et al., 2008), 200 patients with ventilator-associated pneu-
ber of STDs including syphilis, granuloma inguinale (Donovanosis) monia (VAP) received placebo or clarithromycin for three days. Isolated
and chancroid. In the latter two diseases, azithromycin is recommended bacterial pathogens were multidrug-resistant species of Acinetobacter
as first-line therapy (http://www.cdc.gov/std/treatment/2010/toc. baumannii, of Pseudomonas aeruginosa and of Klebsiella pneumoniae,
htm). not included in the antimicrobial spectrum of clarithromycin. The
macrolide significantly reduced by 5.5 days, time until resolution of
6.4. Glomerulonephritis VAP (p = 0.011); by 6.5 days, time until weaning from mechanical
ventilation (p = 0.049); and the relative risk of death by septic shock
Crescentic glomerulonephritis is a severe renal complication of and multiple organ dysfunction (MODS; from 19.00 with placebo to
infectious endocarditis, frequently caused by Gram-positive cocci or 3.78 with clarithromycin; p = 0.048). No difference in incidence of
Gram-negative Bartonella species (Kannan & Mattoo, 2001). The condi- serious adverse events was seen.
tion is not a urinary tract infection per se, but rather derives from im- Immunoparalysis develops during severe sepsis and drives towards
mune activation by bacterial antigens. Bartonella endocarditis immune multiple organ dysfunction syndrome (MODS) and a lethal outcome
complexes are deposited in the glomeruli cause necrotizing crescentic (Boomer et al., 2011). Crucially, clarithromycin treatment was accom-
glomerulonephritis. Interestingly, this form of glomerulonephritis may panied by a reversal of immunoparalysis, most prominently 4 days
improve with antibiotic therapy (Bookman et al., 2004). Resolution of after start of treatment and among patients with septic shock and
glomerulonephritis and normalization of serum creatinine levels was MODS. Compared with placebo, clarithromycin decreased the serum
achieved in several cases by aggressive therapy with azithromycin and IL-10/TNFα ratio; increased monocyte CD86 expression, a marker
a parenteral cephalosporin (ceftriaxone) or by tobramycin/doxycycline of antigen-presentation efficiency; and increased IL-6 production by
combination (Bookman et al., 2004). The efficacy of azithromycin LPS-stimulated monocytes (Spyridaki et al., 2012). Reversal of immuno-
against facultative intracellular Bartonella may be explained by good paralysis was unexpected as attenuation of inflammatory responses in
intracellular penetration of the macrolide (Massei et al., 2005). patients would have been anticipated, based on the results of animal
studies with macrolides (Table 3). It seems that, like azithromycin,
7. Clinical application in sepsis clarithromycin can modulate monocyte function, depending on the
inflammatory environment (see Section 4.1.2). Alternatively, these
Sepsis is an overwhelming inflammatory host response to a bacterial findings, at least in part, may represent an epiphenomenon arising
stimulus. Several animal studies have shown a significant anti- from improvement in the clinical course of sepsis following treatment.
inflammatory effect of azithromycin (Table 3) or clarithromycin in ex- In the second RCT study (Giamarellos-Bourboulis et al., 2013), 600
perimental sepsis. Pretreatment of mice with azithromycin improved patients were enrolled with sepsis due to primary or secondary Gram-
survival and attenuated plasma levels of TNFα in LPS-induced shock negative bacteremia, acute pyelonephritis or acute intrabdominal
(Ivetic Tkalcevic et al., 2006; Tong et al., 2011) and additional immuno- infections, all with microbiologically-proven or clinically-suspected
modulatory effects have been described for clarithromycin. Adminis- Gram-negative bacteria insensitive to clarithromycin. Treatment was
tered i.v. immediately after cecal ligation and puncture in rabbits, extended to four days and the findings of the first RCT were confirmed.
clarithromycin reduced apoptosis in circulating lymphocytes and Mortality due to septic shock and MODS was reduced by clarithromycin
production of TNFα by circulating monocytes, especially within the to 53.6% compared to 73.1% with placebo (p = 0.020). Furthermore,
first 4 h after treatment (Atmatzidis et al., 2011). Pretreatment with among patients with severe sepsis/shock, median time to infection
clarithromycin, but not levofloxacin, also inhibited lung inflammation resolution was 6 days with clarithromycin and 10 days with placebo
and adhesion molecule expression due to high pressure ventilation in (p = 0.037) and was accompanied by a significant shortening of hospi-
mice (Amado-Rodriguez et al., 2013). Similar clinically relevant studies talization time and reduction of costs. Treatment with clarithromycin
on azithromycin have yet to be performed. did not impact the safety of the patients.

Please cite this article as: Parnham, M.J., et al., Azithromycin: Mechanisms of action and their relevance for clinical applications, Pharmacology &
Therapeutics (2014), http://dx.doi.org/10.1016/j.pharmthera.2014.03.003
14 M.J. Parnham et al. / Pharmacology & Therapeutics xxx (2014) xxx–xxx

Current management of sepsis remains largely supportive. Therapy The mechanism of action is considered to be a combination of anti-
targets the underlying infection and aims to support the function of bacterial, immunomodulatory and biofilm-inhibiting effects, associated
failing organs. All agents developed over the last two decades that with the prolonged accumulation of azithromycin in inflamed tissues
inhibit exaggerated pro-inflammatory responses have failed to prolong (Wang, 2010). The improved safety of azithromycin in comparison to
survival or were even harmful. The fact that clarithromycin was able to the current recommended adjuvant antibiotic treatment with amoxi-
reverse the immunoparalytic state arising during severe sepsis, in an cillin and metronidazole is a further added incentive to the introduction
RCT in patients with VAP (Spyridaki et al., 2012), raises hopes that of the macrolide.
macrolides, including azithromycin, may offer a significant advance in Azithromycin is widely used to treat gingival overgrowth, a side-
the management of sepsis. Their efficacy in this condition may help effect of cyclosporin administration. Only 3 RCTs have been carried
explain the benefit of macrolides in patients with CAP. The good efficacy out using a short treatment period, but clear reduction of tissue over-
of azithromycin in severe inflammatory cases of post-transplantation growth was reported in all studies and use of azithromycin for this
bronchiolitis (see Section 5.5), suggests that there may be similarities purpose is recommended (Hirsch et al., 2012).
in the mechanisms involved in effects of macrolides in the two
conditions. 8.3. Ophthamology

8. Other clinical applications Azithromycin is used in ophthalmology for topical treatment of


bacterial conjunctivitis and for off-label use in chronic blepharitis
8.1. Skin disease (inflammation of the eyelid) (Gebre et al., 2012). In the latter, applica-
tion of 1% azithromycin for 2 weeks reduced redness and swelling
Macrolide antibiotics, are widely used for the topical treatment of and improved meibomian gland secretion (Luchs, 2010), probably
acne and rosacea and their use for these and other skin disorders has due to immunomodulatory effects, also observed in a study of
been reviewed (Alzolibani & Zedan, 2012). In several small studies, topical azithromycin in LPS-induced ocular inflammation in the rat
oral azithromycin (for 2–4 weeks) provided significant improvement (Fernandez-Robredo et al., 2013).
in intractable rosacea. In one healthy volunteer-controlled study, stan-
dard dosing with azithromycin for 4 weeks to 17 patients with rosacea, 8.4. Gastric motility
reduced inflammatory score and ameliorated the raised oxidave burst
(chemiluminescence) of leukocytes in biopsies from skin lesions Erythromycin is a stimulator of gastric motility as a consequence of
(Bakar et al., 2007). Since no microbiological analyses were carried agonist activity at motilin receptors (Peeters et al., 1989). Azithromycin
out, it is unclear whether the effects observed were due to anti- similarly stimulates gastric motility and in view of its better safety pro-
inflammatory or antibacterial actions. Currently, rosacea is considered file to that of erythromycin, has been proposed as a treatment for gastric
to be a disease of dysregulated innate immunity. In the absence of motility disorders (Mertens et al., 2009; Rohof et al., 2012). Recently,
RCTs, the place of azithromycin in the treatment of rosacea remains to azithromycin has been shown to have similar potency to erythromycin
be established. as an agonist at human motilin receptors, causing prolonged release of
acetylcholine from human gastric antrum muscle (Broad & Sanger,
8.2. Dentistry 2013).

Over the last 20 years, a number of studies, only a few them RCTs, 9. Safety
have reported beneficial effects of azithromycin in adjuvant treatment
of periodontal disease (Bartold et al., 2013; Hirsch et al., 2012). Micro- A concern about the use of macrolide antibiotics for long-term treat-
biological studies indicate that azithromycin is highly effective against ment of chronic inflammatory diseases, is the possibility of increased
Porphyromonas gingivalis and Aggregatibacter actinomycetemcomitans bacterial resistance, though the relationship between long-term use
(Bartold et al., 2013). and increased resistance is still uncertain (Cameron et al., 2012). A re-
The balance of opinion is that azithromycin relieves infection and cent meta-analysis of six RCTs on the effects of long-term azithromycin
inflammation and reduces pocket depth, particularly in aggressive in chronic lung diseases, revealed that while bacterial colonization de-
forms of periodontal disease, but further controlled studies are needed. creased by 45%, there was a 2.7-fold increase in bacterial resistance (Li

Table 4
Summary of retrospective studied published after 2006 on the impact of macrolide administration on the outcome of community-acquired pneumonia (CAP).

Reference Setting Arms (n patients) Administered macrolide Effect of macrolide intake

(Ambroggio et al., 2012) Hospitalized children β-lactam (15809) NR Reduction of LOS by one day
β-lactam + macrolide (4934) among 31% of children
12–18 y (p b 0.0001)
(Karhu et al., 2013) Severe sepsis due to CAP β-lactam + quinolone (104) Azithromycin or erythromycin No difference
β-lactam + macrolide (106)
(Martin-Loeches et al., 2010) Severe sepsis and MV due to CAP β-lactam + quinolone (54) NR Reduction of risk of death
β-lactam + macrolide (46) (HR: 0.44, p: 0.03)
(Metersky et al., 2007) Bacteremic pneumonia Monotherapy (601) Azithromycin (n = 231) Risk of death after 30 days
Combination therapy (1408) Clarithromycin (n = 65) with macrolide intake 0.61
Erythromycin (n = 32) (p: 0.007)
(Restrepo et al., 2009) Severe sepsis due to CAP Non-macrolide (133) NR Reduction of 90-day mortality
Macrolide (104) (12.5% vs 33.8%, p b 0.0001)
(Tessmer et al., 2009) Hospitalized patients β-lactam (908) NR Independent predictor of lower
β-lactam + macrolide (946) mortality at 14 days
(OR: 0.53, p: 0.031)
(Wilson et al., 2012) Severe sepsis due to CAP β-lactam + quinolone (908) NR Reduction of mean LOS
β-lactam + macrolide (946) (15.9 days vs 24.7 days,
p b 0.0001)

HR: hazard ratio; LOS: length of stay; MV: mechanical ventilation; NR: not reported; OR: odds ratio.

Please cite this article as: Parnham, M.J., et al., Azithromycin: Mechanisms of action and their relevance for clinical applications, Pharmacology &
Therapeutics (2014), http://dx.doi.org/10.1016/j.pharmthera.2014.03.003
M.J. Parnham et al. / Pharmacology & Therapeutics xxx (2014) xxx–xxx 15

et al., 2013). The benefit of long-term macrolide treatment of DPB pa- comparative efficacy with other drugs, azithromycin is only recom-
tients considerably outweighs this concern, particularly since resistance mended for use in DPB, CF, non-CF bronchiectasis and bronchiolitis.
to macrolides is already high in Japan, where DPB is demographically Early, mainly stimulatory effects of azithromycin on immune and
significant (see Section 5.2). In addition, the motilin-agonistic effects epithelial cells are probably mediated by interactions with phospho-
of the antibiotic result in an increased incidence of adverse gastrointes- lipids and Erk1/2. Later modulation of transcription factors AP-1 and
tinal effects on prolonged treatment (Cameron et al., 2012). The devel- NFκB, is accompanied by inhibition of inflammatory cytokine and
opment of novel, non-antibacterial, non-motilin receptor binding, mucin release. Subsequent delayed inhibitory effects on cell function
immunomodulatory macrolides may overcome these concerns are likely related to the high lysosomal accumulation of the macrolide,
(Cameron et al., 2012); Erakovic Haber et al., in press). A further safety with disruption of protein and lipid transport through the Golgi appara-
issue with long-term azithromycin treatment is the possible occurrence tus and ultimate effects on surface receptor expression, including
in a small number of patients of hearing defects (Li et al., 2013). macrophage phenotype changes, and autophagy. These later changes,
For many years, azithromycin has been one of the safest broad particularly on macrophage and dendritic cell phenotype, appear to
spectrum antibiotics. In contrast to other macrolide antibiotics, not underlie many in vivo experimental and clinical immunomodulatory
interacting with CYP3A4 means that it exhibits no metabolic inter- effects of azithromycin, contributing to resolution of acute infections
actions with most other commonly used drugs. It is N20-fold less potent of the airways, genital tract and skin, as well as to reduction of exacerba-
an inhibitor of hERG potassium channels than other macrolide anti- tion frequency in chronic airway diseases.
biotics (Giudicessi & Ackerman, 2013). However, in an observational, Immunomodulation with long-term azithromycin to reduce exacer-
non-randomized study of patients enrolled into a Tennessee Medicaid bations in COPD, CF and non-CF bronchiectasis patients must be
program, patients receiving azithromycin had an incidence of cardio- balanced against the potential for increased bacterial resistance. How-
vascular death 2.88-fold higher than those taking no antibiotic, and ever, a sub-group of post-transplant bronchiolitis patients with pro-
2.49-fold higher than in patients on amoxicillin (Ray et al., 2012). Sub- minent neutrophil infiltration appears to be particularly sensitive to
sequently, the FDA introduced a black box warning about the potential long-term prophylactic therapy with azithromycin, as may be patients
risk of fatal arrhythmias with azithromycin, advising health profes- with severe sepsis treated acutely with the drug. Other promising indi-
sionals to “consider the risk of fatal heart rhythms with azithromycin cations include chronic prostatitis, rosacea and periodontitis.
when considering treatment options for patients who are already at Although well-tolerated with a very good record of safety, azithro-
risk for cardiovascular events.” A further post-marketing safety surveil- mycin has recently been associated with rare cases of torsades des
lance using seven population-based healthcare databases in Denmark, pointes in patients at risk of cardiac arrhymias and polypharmacy with
Italy, and the Netherlands, identified azithromycin, among other medi- azithromycin is generally to be avoided.
cations, as a prime suspect for an association with drug-induced acute
myocardial infarction (Coloma et al., 2013). The US observational Conflict of interest
study was criticized, though, because Medicaid patients generally
have a higher comorbidity than patients in the general population GPP and RV declare that they have no conflicts of interest.
(Giudicessi & Ackerman, 2013). Importantly, a large prospective study
in young to middle-aged Danish people at a low-risk of underlying
Acknowledgments
cardiovascular disease revealed that the risk of cardiovascular death
due to 5-day treatment was 15.4 in comparison to 85.4 per million
MJP has collaborated with Novo Nordisk and serves as an advisor to
azithromycin courses in the Tennessee study (Svanstrom et al., 2013).
Leo Pharma A/S and Xellia Pharmaceuticals ApS. VEH is an employee of
Thus, the risk of a cardiovascular event with azithromycin in otherwise
Fildelta d.o.o. and both MJP and VEH are previous employees of PLIVA
healthy patients is very small, but in patients at risk (i.e. idiopathic or
and GlaxoSmithKline. EGB has received unrestricted educational grants
drug-induced baseline QT prolongation, familial long QT syndrome, or
from Abbott Hellas SA (program 70/3/7447 of the University of Athens).
multiple risk factors) azithromycin should be considered with care
He also serves as an advisor to AbbVie SA.
and polypharmacy should be avoided (Giudicessi & Ackerman, 2013).

References
10. Conclusions
Adriaenssens, N., Coenen, S., Kroes, A. C., Versporten, A., Vankerckhoven, V., Muller, A.,
et al. (2011). European Surveillance of Antimicrobial Consumption (ESAC): systemic
As a broad-spectrum antibacterial, azithromycin shares the same antiviral use in Europe. J Antimicrob Chemother 66, 1897–1905.
mechanism of action as other macrolide antibiotics and its range of Aghai, Z. H., Kode, A., Saslow, J. G., Nakhla, T., Farhath, S., Stahl, G. E., et al. (2007).
Azithromycin suppresses activation of nuclear factor-kappa B and synthesis of pro-
activity is extended through inhibition of bacterial quorum-sensing inflammatory cytokines in tracheal aspirate cells from premature infants. Pediatr
and biofilm. Accumulating more effectively than other macrolides in Res 62, 483–488.
cells, particularly circulating phagocytes, it is delivered in high concen- Albert, R. K., Connett, J., Bailey, W. C., Casaburi, R., Cooper, J. A., Jr., Criner, G. J., et al. (2011).
Azithromycin for prevention of exacerbations of COPD. N Engl J Med 365, 689–698.
trations to sites of infection. This important feature, combined with Allen, N. E. (2002). Effects of macrolide antibiotics on ribosome function. In W. Schönfeld,
the extended plasma half-life of azithromycin, often allows effective & H. A. Kirst (Eds.), Macrolide antibiotics (pp. 261–280). Basel: Birkhäuser Verlag.
single-dose administration for acute bacterial infections. Cellular accu- Altenburg, J., de Graaff, C. S., Stienstra, Y., Sloos, J. H., van Haren, E. H., Koppers, R. J., et al.
(2013). Effect of azithromycin maintenance treatment on infectious exacerbations
mulation, accounting for rapid plasma clearance and extensive tissue among patients with non-cystic fibrosis bronchiectasis: the BAT randomized con-
distribution, offsets the drug's moderate bioavailability. trolled trial. JAMA 309, 1251–1259.
The antibiotic effects of azithromycin are facilitated by its immuno- Alzolibani, A. A., & Zedan, K. (2012). Macrolides in chronic inflammatory skin disorders.
Mediators Inflamm 2012, 159354.
modulatory actions on innate and adaptive immune responses. The
Amado-Rodriguez, L., Gonzalez-Lopez, A., Lopez-Alonso, I., Aguirre, A., Astudillo, A.,
drug appears to exert a biphasic action which may serve to promote Batalla-Solis, E., et al. (2013). Anti-inflammatory effects of clarithromycin in
initial host defence and later reduce bystander tissue injury and ventilator-induced lung injury. Respir Res 14, 52.
Ambroggio, L., Taylor, J. A., Tabb, L. P., Newschaffer, C. J., Evans, A. A., & Shah, S. S. (2012).
promote inflammation resolution. Beneficial effects mediated by
Comparative effectiveness of empiric beta-lactam monotherapy and beta-lactam-
immunomodulation are also seen in a variety of chronic inflammatory macrolide combination therapy in children hospitalized with community-acquired
disorders. These include DPB, post-transplant bronchiolitis (particularly pneumonia. J Pediatr 161, 1097–1103.
severe inflammatory forms) and rosacea. The ability of azithromycin to Amsden, G. W. (1995). Macrolides versus azalides: a drug interaction update. Ann
Pharmacother 29, 906–917.
modulate host response reactions contributes to beneficial effects in Amsden, G. W. (1996). Erythromycin, clarithromycin, and azithromycin: are the differ-
CF, non-CF bronchiectasis, COPD and asthma. Currently, based on ences real? Clin Ther 18, 56–72 (discussion 55).

Please cite this article as: Parnham, M.J., et al., Azithromycin: Mechanisms of action and their relevance for clinical applications, Pharmacology &
Therapeutics (2014), http://dx.doi.org/10.1016/j.pharmthera.2014.03.003
16 M.J. Parnham et al. / Pharmacology & Therapeutics xxx (2014) xxx–xxx

Amsden, G. W. (2001). Advanced-generation macrolides: tissue-directed antibiotics. Int J Branden, E., Koyi, H., Gnarpe, J., Gnarpe, H., & Tornling, G. (2004). Intermittent
Antimicrob Agents 18(Suppl. 1), S11–S15. azithromycin treatment for respiratory symptoms in patients with chronic Chlamydia
Amsden, G. W. (2005). Anti-inflammatory effects of macrolides—an underappreciated pneumoniae infection. Scand J Infect Dis 36, 811–816.
benefit in the treatment of community-acquired respiratory tract infections and Bright, G.M. & Hauske, J.R. (1984). Azahomoerythromycin D derivative and intermediates
chronic inflammatory pulmonary conditions? J Antimicrob Chemother 55, 10–21. thereof. US Patent 4,465,674 A.
Anwar, G. A., Bourke, S. C., Afolabi, G., Middleton, P., Ward, C., & Rutherford, R. M. (2008). Broad, J., & Sanger, G. J. (2013). The antibiotic azithromycin is a motilin receptor agonist in
Effects of long-term low-dose azithromycin in patients with non-CF bronchiectasis. human stomach: comparison with erythromycin. Br J Pharmacol 168, 1859–1867.
Respir Med 102, 1494–1496. Brook, I. (2006). Sinusitis of odontogenic origin. Otolaryngol Head Neck Surg 135, 349–355.
Araki, N., Yanagihara, K., Morinaga, Y., Yamada, K., Nakamura, S., Yamada, Y., et al. (2010). Brusselle, G. G., Vanderstichele, C., Jordens, P., Deman, R., Slabbynck, H., Ringoet, V., et al.
Azithromycin inhibits nontypeable Haemophilus influenzae-induced MUC5AC ex- (2013). Azithromycin for prevention of exacerbations in severe asthma (AZISAST): a
pression and secretion via inhibition of activator protein-1 in human airway epithe- multicentre randomised double-blind placebo-controlled trial. Thorax 68, 322–329.
lial cells. Eur J Pharmacol 644, 209–214. Butorac-Petanjek, B., Parnham, M. J., & Popovic-Grle, S. (2010). Antibiotic therapy for ex-
Asadi, L., Eurich, D. T., Gamble, J. M., Minhas-Sandhu, J. K., Marrie, T. J., & Majumdar, S. R. acerbations of chronic obstructive pulmonary disease (COPD). J Chemother 22,
(2013). Impact of guideline-concordant antibiotics and macrolide/beta-lactam com- 291–297.
binations in 3203 patients hospitalized with pneumonia: prospective cohort study. Cameron, E. J., McSharry, C., Chaudhuri, R., Farrow, S., & Thomson, N. C. (2012). Long-term
Clin Microbiol Infect 19, 257–264. macrolide treatment of chronic inflammatory airway diseases: risks, benefits and fu-
Asadi, L., Sligl, W. I., Eurich, D. T., Colmers, I. N., Tjosvold, L., Marrie, T. J., et al. (2012). ture developments. Clin Exp Allergy 42, 1302–1312.
Macrolide-based regimens and mortality in hospitalized patients with community- Carevic, O., & Djokic, S. (1988). Comparative studies on the effects of erythromycin A and
acquired pneumonia: a systematic review and meta-analysis. Clin Infect Dis 55, azithromycin upon extracellular release of lysosomal enzymes in inflammatory pro-
371–380. cesses. Agents Actions 25, 124–131.
Asakura, E., Nakayama, H., Sugie, M., Zhao, Y. L., Nadai, M., Kitaichi, K., et al. (2004). Carlier, M. B., Garcia-Luque, I., Montenez, J. P., Tulkens, P. M., & Piret, J. (1994). Accumula-
Azithromycin reverses anticancer drug resistance and modifies hepatobiliary excre- tion, release and subcellular localization of azithromycin in phagocytic and non-
tion of doxorubicin in rats. Eur J Pharmacol 484, 333–339. phagocytic cells in culture. Int J Tissue React 16, 211–220.
Asgrimsson, V., Gudjonsson, T., Gudmundsson, G. H., & Baldursson, O. (2006). Novel ef- Champney, W. S., & Burdine, R. (1998). Azithromycin and clarithromycin inhibition of 50S
fects of azithromycin on tight junction proteins in human airway epithelia. ribosomal subunit formation in Staphylococcus aureus cells. Curr Microbiol 36,
Antimicrob Agents Chemother 50, 1805–1812. 119–123.
Atmatzidis, S., Koutelidakis, I., Chatzimavroudis, G., Louis, K., Pistiki, A., Roditis, K., et al. Champney, W. S., Tober, C. L., & Burdine, R. (1998). A comparison of the inhibition of
(2011). Clarithromycin modulates immune responses in experimental peritonitis. translation and 50S ribosomal subunit formation in Staphylococcus aureus cells by
Int J Antimicrob Agents 37, 347–351. nine different macrolide antibiotics. Curr Microbiol 37, 412–417.
Baciu, M., Sebai, S. C., Ces, O., Mulet, X., Clarke, J. A., Shearman, G. C., et al. (2006). Degra- Cigana, C., Nicolis, E., Pasetto, M., Assael, B. M., & Melotti, P. (2006). Anti-inflammatory ef-
dative transport of cationic amphiphilic drugs across phospholipid bilayers. Philos fects of azithromycin in cystic fibrosis airway epithelial cells. Biochem Biophys Res
Transact A Math Phys Eng Sci 364, 2597–2614. Commun 350, 977–982.
Bakar, O., Demircay, Z., Yuksel, M., Haklar, G., & Sanisoglu, Y. (2007). The effect of Clement, A., Tamalet, A., Leroux, E., Ravilly, S., Fauroux, B., & Jais, J. P. (2006). Long term
azithromycin on reactive oxygen species in rosacea. Clin Exp Dermatol 32, effects of azithromycin in patients with cystic fibrosis: a double blind, placebo con-
197–200. trolled trial. Thorax 61, 895–902.
Ballard, H. O., Bernard, P., Qualls, J., Everson, W., & Shook, L. A. (2007). Azithromycin pro- Coates, P., Daniel, R., Houston, A. C., Antrobus, J. H., & Taylor, T. (1991). An open study to
tects against hyperoxic lung injury in neonatal rats. J Investig Med 55, 299–305. compare the pharmacokinetics, safety and tolerability of a multiple-dose regimen of
Ballow, C. H., Amsden, G. W., Highet, V. S., & Forrest, A. (1998). Pharmacokinetics azithromycin in young and elderly volunteers. Eur J Clin Microbiol Infect Dis 10,
of oral azithromycin in serum, urine, polymorphonuclear leucocytes and in- 850–852.
flammatory vs non-inflammatory skin blisters in healthy volunteers. Clin Coloma, P. M., Schuemie, M. J., Trifiro, G., Furlong, L., van Mulligen, E., Bauer-
Drug Investig 15, 159–167. Mehren, A., et al. (2013). Drug-induced acute myocardial infarction: identify-
Banjanac, M., Munic Kos, V., Nujic, K., Vrancic, M., Belamaric, D., Crnkovic, S., et al. (2012). ing ‘prime suspects’ from electronic healthcare records-based surveillance
Anti-inflammatory mechanism of action of azithromycin in LPS-stimulated J774A.1 system. PLoS One 8, e72148.
cells. Pharmacol Res 66, 357–362. Culic, O., Erakovic, V., Cepelak, I., Barisic, K., Brajsa, K., Ferencic, Z., et al. (2002).
Bartold, P. M., du Bois, A. H., Gannon, S., Haynes, D. R., & Hirsch, R. S. (2013). Antibacterial Azithromycin modulates neutrophil function and circulating inflammatory mediators
and immunomodulatory properties of azithromycin treatment implications for peri- in healthy human subjects. Eur J Pharmacol 450, 277–289.
odontitis. Inflammopharmacology 21, 321–338. Culic, O., Erakovic, V., & Parnham, M. J. (2001). Anti-inflammatory effects of macrolide an-
Beigelman, A., Gunsten, S., Mikols, C. L., Vidavsky, I., Cannon, C. L., Brody, S. L., et al. (2009). tibiotics. Eur J Pharmacol 429, 209–229.
Azithromycin attenuates airway inflammation in a noninfectious mouse model of al- Curatolo, W., Liu, P., Johnson, B. A., Hausberger, A., Quan, E., Vendola, T., et al. (2011). Ef-
lergic asthma. Chest 136, 498–506. fects of food on a gastrically degraded drug: azithromycin fast-dissolving gelatin cap-
Beigelman, A., Mikols, C. L., Gunsten, S. P., Cannon, C. L., Brody, S. L., & Walter, M. J. (2010). sules and HPMC capsules. Pharm Res 28, 1531–1539.
Azithromycin attenuates airway inflammation in a mouse model of viral bronchioli- Daenas, C., Hatziefthimiou, A. A., Gourgoulianis, K. I., & Molyvdas, P. A. (2006).
tis. Respir Res 11, 90. Azithromycin has a direct relaxant effect on precontracted airway smooth muscle.
Bermejo-Martin, J. F., Kelvin, D. J., Eiros, J. M., Castrodeza, J., & Ortiz de Lejarazu, R. (2009). Eur J Pharmacol 553, 280–287.
Macrolides for the treatment of severe respiratory illness caused by novel H1N1 Dahl, E. L., & Rosenthal, P. J. (2007). Multiple antibiotics exert delayed effects against
swine influenza viral strains. J Infect Dev Ctries 3, 159–161. the Plasmodium falciparum apicoplast. Antimicrob Agents Chemother 51,
Binder, A. M., Adjemian, J., Olivier, K. N., & Prevots, D. R. (2013). Epidemiology of 3485–3490.
nontuberculous mycobacterial infections and associated chronic macrolide use Dellinger, R. P., Levy, M. M., Rhodes, A., Annane, D., Gerlach, H., Opal, S. M., et al. (2013).
among persons with cystic fibrosis. Am J Respir Crit Care Med 188, 807–812. Surviving sepsis campaign: international guidelines for management of severe sepsis
Blasi, F., Bonardi, D., Aliberti, S., Tarsia, P., Confalonieri, M., Amir, O., et al. (2010). Long- and septic shock: 2012. Crit Care Med 41, 580–637.
term azithromycin use in patients with chronic obstructive pulmonary disease and den Heijer, C. D., van Bijnen, E. M., Paget, W. J., Pringle, M., Goossens, H., Bruggeman, C. A.,
tracheostomy. Pulm Pharmacol Ther 23, 200–207. et al. (2013). Prevalence and resistance of commensal Staphylococcus aureus, includ-
Blondeau, J. M., DeCarolis, E., Metzler, K. L., & Hansen, G. T. (2002). The macrolides. Expert ing meticillin-resistant S. aureus, in nine European countries: a cross-sectional study.
Opin Investig Drugs 11, 189–215. Lancet Infect Dis 13, 409–415.
Bookman, I., Scholey, J. W., Jassal, S. V., Lajoie, G., & Herzenberg, A. M. (2004). Necrotizing Di Paolo, A., Barbara, C., Chella, A., Angeletti, C. A., & Del Tacca, M. (2002). Pharmacokinet-
glomerulonephritis caused by Bartonella henselae endocarditis. Am J Kidney Dis 43, ics of azithromycin in lung tissue, bronchial washing, and plasma in patients given
e25–e30. multiple oral doses of 500 and 1000 mg daily. Pharmacol Res 46, 545–550.
Boomer, J. S., To, K., Chang, K. C., Takasu, O., Osborne, D. F., Walton, A. H., et al. (2011). Im- Dias, R., & Canica, M. (2004). Emergence of invasive erythromycin-resistant Streptococcus
munosuppression in patients who die of sepsis and multiple organ failure. JAMA 306, pneumoniae strains in Portugal: contribution and phylogenetic relatedness of sero-
2594–2605. type 14. J Antimicrob Chemother 54, 1035–1039.
Bosnar, M., Bosnjak, B., Cuzic, S., Hrvacic, B., Marjanovic, N., Glojnaric, I., et al. (2009). Dinos, G. P., Michelinaki, M., & Kalpaxis, D. L. (2001). Insights into the mechanism of
Azithromycin and clarithromycin inhibit lipopolysaccharide-induced murine pulmo- azithromycin interaction with an Escherichia coli functional ribosomal complex. Mol
nary neutrophilia mainly through effects on macrophage-derived granulocyte–mac- Pharmacol 59, 1441–1445.
rophage colony-stimulating factor and interleukin-1beta. J Pharmacol Exp Ther 331, Ehnhage, A., Rautiainen, M., Fang, A. F., & Sanchez, S. P. (2008). Pharmacokinetics of
104–113. azithromycin in serum and sinus fluid after administration of extended-release and
Bosnar, M., Cuzic, S., Bosnjak, B., Nujic, K., Ergovic, G., Marjanovic, N., et al. (2011). immediate-release formulations in patients with acute bacterial sinusitis. Int J
Azithromycin inhibits macrophage interleukin-1beta production through inhibition Antimicrob Agents 31, 561–566.
of activator protein-1 in lipopolysaccharide-induced murine pulmonary neutrophilia. Erakovic Haber, V., Bosnar, M., & Kragol, G. (2014). Design of novel classes of macrolides
Int Immunopharmacol 11, 424–434. for neutrophil dominated inflammatory diseases. Future Medicinal Chemistry 6 in
Bosnar, M., Kelneric, Z., Munic, V., Erakovic, V., & Parnham, M. J. (2005). Cellular uptake press.
and efflux of azithromycin, erythromycin, clarithromycin, telithromycin, and Ernst, E. J., Klepser, M. E., Klepser, T. B., Nightingale, C. H., & Hunsicker, L. G. (2000). Com-
cethromycin. Antimicrob Agents Chemother 49, 2372–2377. parison of the serum and intracellular pharmacokinetics of azithromycin in healthy
Bosnar, M., Kragol, G., Kostrun, S., Vujasinovic, I., Bosnjak, B., Bencetic Mihaljevic, V., and diabetic volunteers. Pharmacotherapy 20, 657–661.
et al. (2012). N′-substituted-2′-O,3′-N-carbonimidoyl bridged macrolides: novel Esposito, S., Bosis, S., Faelli, N., Begliatti, E., Droghetti, R., Tremolati, E., et al. (2005). Role of
anti-inflammatory macrolides without antimicrobial activity. J Med Chem 55, atypical bacteria and azithromycin therapy for children with recurrent respiratory
6111–6123. tract infections. Pediatr Infect Dis J 24, 438–444.

Please cite this article as: Parnham, M.J., et al., Azithromycin: Mechanisms of action and their relevance for clinical applications, Pharmacology &
Therapeutics (2014), http://dx.doi.org/10.1016/j.pharmthera.2014.03.003
M.J. Parnham et al. / Pharmacology & Therapeutics xxx (2014) xxx–xxx 17

Fa, N., Lins, L., Courtoy, P. J., Dufrene, Y., Van Der Smissen, P., Brasseur, R., et al. (2007). De- Gladue, R. P., & Snider, M. E. (1990). Intracellular accumulation of azithromycin by cul-
crease of elastic moduli of DOPC bilayers induced by a macrolide antibiotic, tured human fibroblasts. Antimicrob Agents Chemother 34, 1056–1060.
azithromycin. Biochim Biophys Acta 1768, 1830–1838. Glojnaric, I., Cuzic, S., Erakovic-Haber, V., & Parnham, M. J. (2007). The serum amyloid A
Falk, L., Fredlund, H., & Jensen, J. S. (2003). Tetracycline treatment does not eradicate My- response to sterile silver nitrate in mice and its inhibition by dexamethasone and
coplasma genitalium. Sex Transm Infect 79, 318–319. macrolide antibiotics. Int Immunopharmacol 7, 1544–1551.
Fang, A. F., Palmer, J. N., Chiu, A. G., Blumer, J. L., Crownover, P. H., Campbell, M. D., et al. Gomi, K., Yashima, A., Iino, F., Kanazashi, M., Nagano, T., Shibukawa, N., et al. (2007). Drug
(2009). Pharmacokinetics of azithromycin in plasma and sinus mucosal tissue follow- concentration in inflamed periodontal tissues after systemically administered
ing administration of extended-release or immediate-release formulations in adult azithromycin. J Periodontol 78, 918–923.
patients with chronic rhinosinusitis. Int J Antimicrob Agents 34, 67–71. Gottlieb, J., Szangolies, J., Koehnlein, T., Golpon, H., Simon, A., & Welte, T. (2008). Long-
Farrell, D. J., & Jenkins, S. G. (2004). Distribution across the USA of macrolide resistance term azithromycin for bronchiolitis obliterans syndrome after lung transplantation.
and macrolide resistance mechanisms among Streptococcus pneumoniae isolates col- Transplantation 85, 36–41.
lected from patients with respiratory tract infections: PROTEKT US 2001–2002. J Halldorsson, S., Gudjonsson, T., Gottfredsson, M., Singh, P. K., Gudmundsson, G. H., &
Antimicrob Chemother 54(Suppl. 1), i17–i22. Baldursson, O. (2010). Azithromycin maintains airway epithelial integrity during
Federica, M., Nadia, S., Monica, M., Alessandro, C., Tiberio, O., Francesco, B., et al. (2011). Pseudomonas aeruginosa infection. Am J Respir Cell Mol Biol 42, 62–68.
Clinical and immunological evaluation of 12-month azithromycin therapy in chronic Hand, W. L., & Hand, D. L. (2001). Characteristics and mechanisms of azithromycin accu-
lung allograft rejection. Clin Transplant 25, E381–E389. mulation and efflux in human polymorphonuclear leukocytes. Int J Antimicrob Agents
Felmingham, D., & Gruneberg, R. N. (2000). The Alexander Project 1996–1997: latest 18, 419–425.
susceptibility data from this international study of bacterial pathogens from Hansen, J. L., Ippolito, J. A., Ban, N., Nissen, P., Moore, P. B., & Steitz, T. A. (2002). The struc-
community-acquired lower respiratory tract infections. J Antimicrob Chemother tures of four macrolide antibiotics bound to the large ribosomal subunit. Mol Cell 10,
45, 191–203. 117–128.
Feola, D. J., Garvy, B. A., Cory, T. J., Birket, S. E., Hoy, H., Hayes, D., Jr., et al. (2010). Hao, K., Qi, Q., Hao, H., Wang, G., Chen, Y., Liang, Y., et al. (2013). The pharmacokinetic–
Azithromycin alters macrophage phenotype and pulmonary compartmentalization pharmacodynamic model of azithromycin for lipopolysaccharide-induced
during lung infection with Pseudomonas. Antimicrob Agents Chemother 54, depressive-like behavior in mice. PLoS One 8, e54981.
2437–2447. Hassan, H. E., Othman, A. A., Eddington, N. D., Duffy, L., Xiao, L., Waites, K. B., et al. (2011).
Fernandes, A. C., Anderson, R., Theron, A. J., Joone, G., & Van Rensburg, C. E. (1984). En- Pharmacokinetics, safety, and biologic effects of azithromycin in extremely preterm
hancement of human polymorphonuclear leucocyte motility by erythromycin infants at risk for ureaplasma colonization and bronchopulmonary dysplasia. J Clin
in vitro and in vivo. S Afr Med J 66, 173–177. Pharmacol 51, 1264–1275.
Fernandez-Robredo, P., Recalde, S., Moreno-Orduna, M., Garcia-Garcia, L., Zarranz- Healy, D. P., Silverman, P. A., Neely, A. N., Holder, I. A., & Babcock, G. E. (2002). Effect of
Ventura, J., & Garcia-Layana, A. (2013). Azithromycin reduces inflammation in a rat antibiotics on polymorphonuclear neutrophil apoptosis. Pharmacotherapy 22,
model of acute conjunctivitis. Mol Vis 19, 153–165. 578–585.
Fernandez, A. D., Elmore, M. K., & Metzger, D. W. (2004). Azithromycin modulates murine Henry, D. C., Riffer, E., Sokol, W. N., Chaudry, N. I., & Swanson, R. N. (2003). Randomized
immune responses to pneumococcal conjugate vaccine and inhibits nasal clearance double-blind study comparing 3- and 6-day regimens of azithromycin with a 10-
of bacteria. J Infect Dis 190, 1762–1766. day amoxicillin-clavulanate regimen for treatment of acute bacterial sinusitis.
Fimia, G. M., & Piacentini, M. (2010). Regulation of autophagy in mammals and its inter- Antimicrob Agents Chemother 47, 2770–2774.
play with apoptosis. Cell Mol Life Sci 67, 1581–1588. Hirsch, R., Deng, H., & Laohachai, M. N. (2012). Azithromycin in periodontal treatment:
Fleet, J. L., Shariff, S. Z., Bailey, D. G., Gandhi, S., Juurlink, D. N., Nash, D. M., et al. (2013). more than an antibiotic. J Periodontal Res 47, 137–148.
Comparing two types of macrolide antibiotics for the purpose of assessing Hodge, S., Hodge, G., Brozyna, S., Jersmann, H., Holmes, M., & Reynolds, P. N. (2006).
population-based drug interactions. BMJ Open 3. Azithromycin increases phagocytosis of apoptotic bronchial epithelial cells by alveo-
Florey, O., & Overholtzer, M. (2012). Autophagy proteins in macroendocytic engulfment. lar macrophages. Eur Respir J 28, 486–495.
Trends Cell Biol 22, 374–380. Hodge, S., Hodge, G., Jersmann, H., Matthews, G., Ahern, J., Holmes, M., et al. (2008).
Foulds, G., Madsen, P., Cox, C., Shepard, R., & Johnson, R. (1991). Concentration of Azithromycin improves macrophage phagocytic function and expression of mannose
azithromycin in human prostatic tissue. Eur J Clin Microbiol Infect Dis 10, 868–871. receptor in chronic obstructive pulmonary disease. Am J Respir Crit Care Med 178,
Foulds, G., Shepard, R. M., & Johnson, R. B. (1990). The pharmacokinetics of azithromycin 139–148.
in human serum and tissues. J Antimicrob Chemother 25(Suppl. A), 73–82. Hodge, S., & Reynolds, P. N. (2012). Low-dose azithromycin improves phagocytosis of
Franceschi, F., Kanyo, Z., Sherer, E. C., & Sutcliffe, J. (2004). Macrolide resistance from the bacteria by both alveolar and monocyte-derived macrophages in chronic obstructive
ribosome perspective. Curr Drug Targets Infect Disord 4, 177–191. pulmonary disease subjects. Respirology 17, 802–807.
Garver, E., Hugger, E. D., Shearn, S. P., Rao, A., Dawson, P. A., Davis, C. B., et al. (2008). In- Hoffler, D., Koeppe, P., & Paeske, B. (1995). Pharmacokinetics of azithromycin in normal
volvement of intestinal uptake transporters in the absorption of azithromycin and and impaired renal function. Infection 23, 356–361.
clarithromycin in the rat. Drug Metab Dispos 36, 2492–2498. Hoffmann, N., Lee, B., Hentzer, M., Rasmussen, T. B., Song, Z., Johansen, H. K., et al. (2007).
Gebre, T., Ayele, B., Zerihun, M., Genet, A., Stoller, N. E., Zhou, Z., et al. (2012). Comparison Azithromycin blocks quorum sensing and alginate polymer formation and increases
of annual versus twice-yearly mass azithromycin treatment for hyperendemic tra- the sensitivity to serum and stationary-growth-phase killing of Pseudomonas
choma in Ethiopia: a cluster-randomised trial. Lancet 379, 143–151. aeruginosa and attenuates chronic P. aeruginosa lung infection in Cftr(−/−) mice.
Gerbaux, C., Van Bambeke, F., Montenez, J. P., Piret, J., Morlighem, G., & Tulkens, P. M. Antimicrob Agents Chemother 51, 3677–3687.
(1996). Hyperactivity of cathepsin B and other lysosomal enzymes in fibroblasts ex- Hoiby, N. (2011). Recent advances in the treatment of Pseudomonas aeruginosa infections
posed to azithromycin, a dicationic macrolide antibiotic with exceptional tissue accu- in cystic fibrosis. BMC Med 9, 32.
mulation. FEBS Lett 394, 307–310. Ianaro, A., Ialenti, A., Maffia, P., Sautebin, L., Rombola, L., Carnuccio, R., et al. (2000).
Gerhardt, S. G., McDyer, J. F., Girgis, R. E., Conte, J. V., Yang, S. C., & Orens, J. B. (2003). Main- Anti-inflammatory activity of macrolide antibiotics. J Pharmacol Exp Ther 292,
tenance azithromycin therapy for bronchiolitis obliterans syndrome: results of a pilot 156–163.
study. Am J Respir Crit Care Med 168, 121–125. Ikeda, K., Hirayama, M., Hirota, Y., Asa, E., Seki, J., & Tanaka, Y. (2008). Drug-induced
Geudens, N., Timmermans, L., Vanhooren, H., Vanaudenaerde, B. M., Vos, R., Van De phospholipidosis is caused by blockade of mannose 6-phosphate receptor-mediated
Wauwer, C., et al. (2008). Azithromycin reduces airway inflammation in a murine targeting of lysosomal enzymes. Biochem Biophys Res Commun 377, 268–274.
model of lung ischaemia reperfusion injury. Transpl Int 21, 688–695. Ikegaya, S., Inai, K., Iwasaki, H., Naiki, H., & Ueda, T. (2009). Azithromycin reduces tumor
Giamarellos-Bourboulis, E. J., Mylona, V., Antonopoulou, A., Tsangaris, I., Koutelidakis, I., necrosis factor-alpha production in lipopolysaccharide-stimulated THP-1 monocytic
Marioli, A., et al. (2013). Effect of clarithromycin in patients with suspected Gram- cells by modification of stress response and p38 MAPK pathway. J Chemother 21,
negative sepsis: results of a randomized controlled trial. J Antimicrob Chemother, 396–402.
http://dx.doi.org/10.1093/jac/dkt475. Imamura, Y., Yanagihara, K., Mizuta, Y., Seki, M., Ohno, H., Higashiyama, Y., et al. (2004).
Giamarellos-Bourboulis, E. J., Pechere, J. C., Routsi, C., Plachouras, D., Kollias, S., Azithromycin inhibits MUC5AC production induced by the Pseudomonas aeruginosa
Raftogiannis, M., et al. (2008). Effect of clarithromycin in patients with sepsis and autoinducer N-(3-Oxododecanoyl) homoserine lactone in NCI-H292 Cells.
ventilator-associated pneumonia. Clin Infect Dis 46, 1157–1164. Antimicrob Agents Chemother 48, 3457–3461.
Giannopoulos, A., Koratzanis, G., Giamarellos-Bourboulis, E. J., Panou, C., Adamakis, I., & Inoue, M., Farrell, D. J., Kaneko, K., Akizawa, K., Fujita, S., Kaku, M., et al. (2008). Antimicro-
Giamarellou, H. (2001). Pharmacokinetics of clarithromycin in the prostate: implica- bial susceptibility of respiratory tract pathogens in Japan during PROTEKT years 1–5
tions for the treatment of chronic abacterial prostatitis. J Urol 165, 97–99. (1999–2004). Microb Drug Resist 14, 109–117.
Gielen, V., Johnston, S. L., & Edwards, M. R. (2010). Azithromycin induces anti-viral re- Ishimatsu, Y., Kadota, J., Iwashita, T., Nagata, T., Ishii, H., Shikuwa, C., et al. (2004).
sponses in bronchial epithelial cells. Eur Respir J 36, 646–654. Macrolide antibiotics induce apoptosis of human peripheral lymphocytes in vitro.
Giovanetti, E., Brenciani, A., Burioni, R., & Varaldo, P. E. (2002). A novel efflux system in Int J Antimicrob Agents 24, 247–253.
inducibly erythromycin-resistant strains of Streptococcus pyogenes. Antimicrob Ishimoto, H., Mukae, H., Sakamoto, N., Amenomori, M., Kitazaki, T., Imamura, Y., et al.
Agents Chemother 46, 3750–3755. (2009). Different effects of telithromycin on MUC5AC production induced by
Giovanetti, E., Montanari, M. P., Mingoia, M., & Varaldo, P. E. (1999). Phenotypes and human neutrophil peptide-1 or lipopolysaccharide in NCI-H292 cells compared
genotypes of erythromycin-resistant Streptococcus pyogenes strains in Italy and with azithromycin and clarithromycin. J Antimicrob Chemother 63, 109–114.
heterogeneity of inducibly resistant strains. Antimicrob Agents Chemother 43, Ivetic Tkalcevic, V., Bosnjak, B., Hrvacic, B., Bosnar, M., Marjanovic, N., Ferencic, Z., et al.
1935–1940. (2006). Anti-inflammatory activity of azithromycin attenuates the effects of lipopoly-
Giudicessi, J. R., & Ackerman, M. J. (2013). Azithromycin and risk of sudden cardiac death: saccharide administration in mice. Eur J Pharmacol 539, 131–138.
guilty as charged or falsely accused? Cleve Clin J Med 80, 539–544. Ivetic Tkalcevic, V., Cuzic, S., Kramaric, M. D., Parnham, M. J., & Erakovic Haber, V. (2012).
Gladue, R. P., Bright, G. M., Isaacson, R. E., & Newborg, M. F. (1989). In vitro and in vivo Topical azithromycin and clarithromycin inhibit acute and chronic skin inflammation
uptake of azithromycin (CP-62,993) by phagocytic cells: possible mechanism of de- in sensitized mice, with apparent selectivity for Th2-mediated processes in delayed-
livery and release at sites of infection. Antimicrob Agents Chemother 33, 277–282. type hypersensitivity. Inflammation 35, 192–205.

Please cite this article as: Parnham, M.J., et al., Azithromycin: Mechanisms of action and their relevance for clinical applications, Pharmacology &
Therapeutics (2014), http://dx.doi.org/10.1016/j.pharmthera.2014.03.003
18 M.J. Parnham et al. / Pharmacology & Therapeutics xxx (2014) xxx–xxx

Iwamoto, S., Azuma, E., Kumamoto, T., Hirayama, M., Yoshida, T., Ito, M., et al. (2013). Ef- Li, H., Liu, D. H., Chen, L. L., Zhao, Q., Yu, Y. Z., Ding, J. J., et al. (2013). Adverse effects of
ficacy of azithromycin in preventing lethal graft-versus-host disease. Clin Exp long-term azithromycin use in patients with chronic lung diseases: a meta-analysis.
Immunol 171, 338–345. Antimicrob Agents Chemother 58, 511–517.
Iwamoto, S., Kumamoto, T., Azuma, E., Hirayama, M., Ito, M., Amano, K., et al. (2011). The Li, H., Zhou, Y., Fan, F., Zhang, Y., Li, X., Yu, H., et al. (2011). Effect of azithromycin on pa-
effect of azithromycin on the maturation and function of murine bone marrow- tients with diffuse panbronchiolitis: retrospective study of 51 cases. Intern Med 50,
derived dendritic cells. Clin Exp Immunol 166, 385–392. 1663–1669.
Jacobs, M. R., & Johnson, C. E. (2003). Macrolide resistance: an increasing concern for Lin, H. C., Wang, C. H., Liu, C. Y., Yu, C. T., & Kuo, H. P. (2000). Erythromycin inhibits beta2-
treatment failure in children. Pediatr Infect Dis J 22, S131–S138. integrins (CD11b/CD18) expression, interleukin-8 release and intracellular oxidative
Jacobs, R. F., Maples, H. D., Aranda, J. V., Espinoza, G. M., Knirsch, C., Chandra, R., et al. metabolism in neutrophils. Respir Med 94, 654–660.
(2005). Pharmacokinetics of intravenously administered azithromycin in pediatric Lin, S. J., Yan, D. C., Lee, W. I., Kuo, M. L., Hsiao, H. S., & Lee, P. Y. (2012). Effect of
patients. Pediatr Infect Dis J 24, 34–39. azithromycin on natural killer cell function. Int Immunopharmacol 13, 8–14.
Jenkins, S. G., Brown, S. D., & Farrell, D. J. (2008). Trends in antibacterial resistance among Liu, P., Fang, A. F., LaBadie, R. R., Crownover, P. H., & Arguedas, A. G. (2011). Comparison of
Streptococcus pneumoniae isolated in the USA: update from PROTEKT US Years 1–4. azithromycin pharmacokinetics following single oral doses of extended-release and
Ann Clin Microbiol Antimicrob 7, 1. immediate-release formulations in children with acute otitis media. Antimicrob
Ju, J. S., & Weihl, C. C. (2010). Inclusion body myopathy, Paget's disease of the bone and Agents Chemother 55, 5022–5026.
fronto-temporal dementia: a disorder of autophagy. Hum Mol Genet 19, R38–R45. Lode, H., Borner, K., Koeppe, P., & Schaberg, T. (1996). Azithromycin—review of key chemical,
Kannan, S., & Mattoo, T. K. (2001). Diffuse crescentic glomerulonephritis in bacterial en- pharmacokinetic and microbiological features. J Antimicrob Chemother 37(Suppl. C), 1–8.
docarditis. Pediatr Nephrol 16, 423–428. Lucchi, M., Damle, B., Fang, A., de Caprariis, P. J., Mussi, A., Sanchez, S. P., et al. (2008).
Kanoh, S., & Rubin, B. K. (2010). Mechanisms of action and clinical application of Pharmacokinetics of azithromycin in serum, bronchial washings, alveolar macro-
macrolides as immunomodulatory medications. Clin Microbiol Rev 23, 590–615. phages and lung tissue following a single oral dose of extended or immediate release
Karhu, J., Ala-Kokko, T. I., Ohtonen, P., & Syrjala, H. (2013). Severe community-acquired formulations of azithromycin. J Antimicrob Chemother 61, 884–891.
pneumonia treated with beta-lactam-respiratory quinolone vs. beta-lactam- Luchs, J. (2010). Azithromycin in DuraSite for the treatment of blepharitis. Clin
macrolide combination. Acta Anaesthesiol Scand 57, 587–593. Ophthalmol 4, 681–688.
Karpov, O. I. (1999). The clinical and economic efficacies of short courses of azithromycin Luke, D. R., & Foulds, G. (1997). Disposition of oral azithromycin in humans. Clin
in acute sinusitis. Antibiot Khimioter 44, 28–32. Pharmacol Ther 61, 641–648.
Kent, J. R., Almond, M. K., & Dhillon, S. (2001). Azithromycin: an assessment of its phar- Luo, L., Perelman, J. M., Kolosov, V. P., & Zhou, X. D. (2011). Inhibition of airway mucous
macokinetics and therapeutic potential in CAPD. Perit Dial Int 21, 372–377. hypersecretion by azithromycin through matrix metalloproteinase 9. Zhonghua Yi
Kezerashvili, A., Khattak, H., Barsky, A., Nazari, R., & Fisher, J. D. (2007). Azithromycin as a Xue Za Zhi 91, 689–693.
cause of QT-interval prolongation and torsade de pointes in the absence of other Lutz, L., Pereira, D. C., Paiva, R. M., Zavascki, A. P., & Barth, A. L. (2012). Macrolides
known precipitating factors. J Interv Card Electrophysiol 18, 243–246. decrease the minimal inhibitory concentration of anti-pseudomonal agents
Khair, O. A., Devalia, J. L., Abdelaziz, M. M., Sapsford, R. J., & Davies, R. J. (1995). Effect of against Pseudomonas aeruginosa from cystic fibrosis patients in biofilm. BMC
erythromycin on Haemophilus influenzae endotoxin-induced release of IL-6, IL-8 Microbiol 12, 196.
and sICAM-1 by cultured human bronchial epithelial cells. Eur Respir J 8, 1451–1457. Magri, V., Marras, E., Skerk, V., Markotic, A., Restelli, A., Garlaschi, M. C., et al. (2010). Erad-
Khan, A. A., Slifer, T. R., Araujo, F. G., & Remington, J. S. (1999). Effect of clarithromycin and ication of Chlamydia trachomatis parallels symptom regression in chronic bacterial
azithromycin on production of cytokines by human monocytes. Int J Antimicrob prostatitis patients treated with a fluoroquinolone-macrolide combination.
Agents 11, 121–132. Andrologia 42, 366–375.
Kita, E., Sawaki, M., Nishikawa, F., Mikasa, K., Yagyu, Y., Takeuchi, S., et al. (1990). En- Magri, V., Trinchieri, A., Pozzi, G., Restelli, A., Garlaschi, M. C., Torresani, E., et al. (2007).
hanced interleukin production after long-term administration of erythromycin stea- Efficacy of repeated cycles of combination therapy for the eradication of infecting or-
rate. Pharmacology 41, 177–183. ganisms in chronic bacterial prostatitis. Int J Antimicrob Agents 29, 549–556.
Kneyber, M. C., van Woensel, J. B., Uijtendaal, E., Uiterwaal, C. S., & Kimpen, J. L.Dutch Malhotra-Kumar, S., Lammens, C., Coenen, S., Van Herck, K., & Goossens, H. (2007). Effect
Antibiotics in, R. S. V. T. R. G. (2008). Azithromycin does not improve disease course of azithromycin and clarithromycin therapy on pharyngeal carriage of macrolide-
in hospitalized infants with respiratory syncytial virus (RSV) lower respiratory tract resistant streptococci in healthy volunteers: a randomised, double-blind, placebo-
disease: a randomized equivalence trial. Pediatr Pulmonol 43, 142–149. controlled study. Lancet 369, 482–490.
Kobayashi, H., Ohgaki, N., & Takeda, H. (1993). Therapeutic possibilities for diffuse Mann, J. M., Sha, K. K., Kline, G., Breuer, F. U., & Miller, A. (2005). World Trade Center dys-
panbronchiolitis. Int J Antimicrob Agents 3(Suppl. 1), S81–S86. pnea: bronchiolitis obliterans with functional improvement: a case report. Am J Ind
Kobayashi, Y., Wada, H., Rossios, C., Takagi, D., Higaki, M., Mikura, S., et al. (2013). A novel Med 48, 225–229.
macrolide solithromycin exerts superior anti-inflammatory effect via NF-kappaB in- Marjanovic, N., Bosnar, M., Michielin, F., Wille, D. R., Anic-Milic, T., Culic, O., et al. (2011).
hibition. J Pharmacol Exp Ther 345, 76–84. Macrolide antibiotics broadly and distinctively inhibit cytokine and chemokine pro-
Kobrehel, G., Radobolja, G., Tamburasev, Z., Djokic, S. (1982). 11-Aza-10-deoxo-10- duction by COPD sputum cells in vitro. Pharmacol Res 63, 389–397.
dihydroerythromycin A and derivatives thereof as well as a process for their prepara- Marple, B. F., Roberts, C. S., Frytak, J. R., Schabert, V. F., Wegner, J. C., Bhattacharyya, H.,
tion. U.S. Patent 4,328,334. et al. (2010). Azithromycin extended release vs amoxicillin/clavulanate: symptom
Koch, C. C., Esteban, D. J., Chin, A. C., Olson, M. E., Read, R. R., Ceri, H., et al. (2000). Apoptosis, resolution in acute sinusitis. Am J Otolaryngol 31, 1–8.
oxidative metabolism and interleukin-8 production in human neutrophils exposed to Martin-Loeches, I., Lisboa, T., Rodriguez, A., Putensen, C., Annane, D., Garnacho-Montero, J.,
azithromycin: effects of Streptococcus pneumoniae. J Antimicrob Chemother 46, 19–26. et al. (2010). Combination antibiotic therapy with macrolides improves survival in
Kohler, T., Dumas, J. L., & Van Delden, C. (2007). Ribosome protection prevents intubated patients with community-acquired pneumonia. Intensive Care Med 36,
azithromycin-mediated quorum-sensing modulation and stationary-phase killing of 612–620.
Pseudomonas aeruginosa. Antimicrob Agents Chemother 51, 4243–4248. Massei, F., Gori, L., Macchia, P., & Maggiore, G. (2005). The expanded spectrum of
Kolumbic Lakos, A., Skerk, V., Malekovic, G., Dujnic Spoljarevic, T., Kovacic, D., Pasini, M., bartonellosis in children. Infect Dis Clin North Am 19, 691–711.
et al. (2011). A switch therapy protocol with intravenous azithromycin and ciproflox- Matijasic, M., Munic Kos, V., Nujic, K., Cuzic, S., Padovan, J., Kragol, G., et al. (2012). Fluo-
acin combination for severe, relapsing chronic bacterial prostatitis: a prospective rescently labeled macrolides as a tool for monitoring cellular and tissue distribution
non-comparative pilot study. J Chemother 23, 350–353. of azithromycin. Pharmacol Res 66, 332–342.
Konno, S., Adachi, M., Asano, K., Kawazoe, T., Okamoto, K., & Takahashi, T. (1992). Influ- Matsumura, Y., Mitani, A., Suga, T., Kamiya, Y., Kikuchi, T., Tanaka, S., et al. (2011).
ences of roxithromycin on cell-mediated immune responses. Life Sci 51, L107–L112. Azithromycin may inhibit interleukin-8 through suppression of Rac1 and a nuclear
Kornfeld, S. (1992). Structure and function of the mannose 6-phosphate/insulinlike factor-kappa B pathway in KB cells stimulated with lipopolysaccharide. J Periodontol
growth factor II receptors. Annu Rev Biochem 61, 307–330. 82, 1623–1631.
Kosol, S., Schrank, E., Krajacic, M. B., Wagner, G. E., Meyer, N. H., Gobl, C., et al. (2012). Matzneller, P., Krasniqi, S., Kinzig, M., Sorgel, F., Huttner, S., Lackner, E., et al. (2013).
Probing the interactions of macrolide antibiotics with membrane-mimetics by NMR Blood, tissue, and intracellular concentrations of azithromycin during and after end
spectroscopy. J Med Chem 55, 5632–5636. of therapy. Antimicrob Agents Chemother 57, 1736–1742.
Kotsaki, A., & Giamarellos-Bourboulis, E. J. (2012). Emerging drugs for the treatment of Mazzei, T., Surrenti, C., Novelli, A., Crispo, A., Fallani, S., Carla, V., et al. (1993). Pharmaco-
sepsis. Expert Opin Emerg Drugs 17, 379–391. kinetics of azithromycin in patients with impaired hepatic function. J Antimicrob
Kudoh, S. (2004). Applying lessons learned in the treatment of diffuse panbronchiolitis to Chemother 31(Suppl. E), 57–63.
other chronic inflammatory diseases. Am J Med 117(Suppl. 9A), 12S–19S. McDonald, P. J., & Pruul, H. (1991). Phagocyte uptake and transport of azithromycin. Eur J
Kudoh, S., Azuma, A., Yamamoto, M., Izumi, T., & Ando, M. (1998). Improvement of surviv- Clin Microbiol Infect Dis 10, 828–833.
al in patients with diffuse panbronchiolitis treated with low-dose erythromycin. Am J Mena, L. A., Mroczkowski, T. F., Nsuami, M., & Martin, D. H. (2009). A randomized compar-
Respir Crit Care Med 157, 1829–1832. ison of azithromycin and doxycycline for the treatment of Mycoplasma genitalium-
Lakos, A. K., Pangercic, A., Gasparic, M., Kukuruzovic, M. M., Kovacic, D., & Barsic, B. (2012). positive urethritis in men. Clin Infect Dis 48, 1649–1654.
Safety and effectiveness of azithromycin in the treatment of respiratory infections in Mencarelli, A., Distrutti, E., Renga, B., Cipriani, S., Palladino, G., Booth, C., et al. (2011). De-
children. Curr Med Res Opin 28, 155–162. velopment of non-antibiotic macrolide that corrects inflammation-driven immune
Lau, C. Y., & Qureshi, A. K. (2002). Azithromycin versus doxycycline for genital dysfunction in models of inflammatory bowel diseases and arthritis. Eur J Pharmacol
chlamydial infections: a meta-analysis of randomized clinical trials. Sex 665, 29–39.
Transm Dis 29, 497–502. Mergenhagen, K. A., Olbrych, P. M., Mattappallil, A., Krajewski, M. P., & Ott, M. C. (2013).
Leclercq, R. (2002). Mechanisms of resistance to macrolides and lincosamides: nature of Effect of azithromycin on anticoagulation-related outcomes in geriatric patients re-
the resistance elements and their clinical implications. Clin Infect Dis 34, 482–492. ceiving warfarin. Clin Ther 35, 425–430.
Legssyer, R., Huaux, F., Lebacq, J., Delos, M., Marbaix, E., Lebecque, P., et al. (2006). Mertens, V., Blondeau, K., Pauwels, A., Farre, R., Vanaudenaerde, B., Vos, R., et al. (2009).
Azithromycin reduces spontaneous and induced inflammation in DeltaF508 cystic fi- Azithromycin reduces gastroesophageal reflux and aspiration in lung transplant re-
brosis mice. Respir Res 7, 134. cipients. Dig Dis Sci 54, 972–979.

Please cite this article as: Parnham, M.J., et al., Azithromycin: Mechanisms of action and their relevance for clinical applications, Pharmacology &
Therapeutics (2014), http://dx.doi.org/10.1016/j.pharmthera.2014.03.003
M.J. Parnham et al. / Pharmacology & Therapeutics xxx (2014) xxx–xxx 19

Metersky, M. L., Ma, A., Houck, P. M., & Bratzler, D. W. (2007). Antibiotics for bacteremic Parnham, M. J. (2005). Antibiotics, inflammation and its resolution: an overview. In B. K.
pneumonia: improved outcomes with macrolides but not fluoroquinolones. Chest Rubin, & J. Tamaoka (Eds.), Antibiotics as Anti-Inflammatory and Immunomodulatory
131, 466–473. Agents (pp. 27–47). Basel: Birkhäuser Verlag.
Meyer, M., Huaux, F., Gavilanes, X., van den Brule, S., Lebecque, P., Lo Re, S., et al. (2009). Parnham, M. J., Culic, O., Erakovic, V., Munic, V., Popovic-Grle, S., Barisic, K., et al. (2005).
Azithromycin reduces exaggerated cytokine production by M1 alveolar macrophages Modulation of neutrophil and inflammation markers in chronic obstructive pulmo-
in cystic fibrosis. Am J Respir Cell Mol Biol 41, 590–602. nary disease by short-term azithromycin treatment. Eur J Pharmacol 517, 132–143.
Milberg, P., Eckardt, L., Bruns, H. J., Biertz, J., Ramtin, S., Reinsch, N., et al. (2002). Divergent Pascual, A., Conejo, M. C., Garcia, I., & Perea, E. J. (1995). Factors affecting the intracellular
proarrhythmic potential of macrolide antibiotics despite similar QT prolongation: fast accumulation and activity of azithromycin. J Antimicrob Chemother 35, 85–93.
phase 3 repolarization prevents early afterdepolarizations and torsade de pointes. J Peeters, T., Matthijs, G., Depoortere, I., Cachet, T., Hoogmartens, J., & Vantrappen, G.
Pharmacol Exp Ther 303, 218–225. (1989). Erythromycin is a motilin receptor agonist. Am J Physiol 257, G470–G474.
Miller, S. A., Selzman, C. H., Shames, B. D., Barton, H. A., Johnson, S. M., & Harken, A. H. Pene Dumitrescu, T., Anic-Milic, T., Oreskovic, K., Padovan, J., Brouwer, K. L., Zuo, P., et al.
(2000). Chlamydia pneumoniae activates nuclear factor kappaB and activator protein (2013). Development of a population pharmacokinetic model to describe
1 in human vascular smooth muscle and induces cellular proliferation. J Surg Res 90, azithromycin whole-blood and plasma concentrations over time in healthy subjects.
76–81. Antimicrob Agents Chemother 57, 3194–3201.
Millrose, M., Kruse, M., Flick, B., & Stahlmann, R. (2009). Effects of macrolides on proin- Perletti, G., Skerk, V., Magri, V., Markotic, A., Mazzoli, S., Parnham, M. J., et al. (2011).
flammatory epitopes on endothelial cells in vitro. Arch Toxicol 83, 469–476. Macrolides for the treatment of chronic bacterial prostatitis: an effective application
Min, J. Y., & Jang, Y. J. (2012). Macrolide therapy in respiratory viral infections. Mediators of their unique pharmacokinetic and pharmacodynamic profile (Review). Mol Med
Inflamm 2012, 649570. Rep 4, 1035–1044.
Miossec-Bartoli, C., Pilatre, L., Peyron, P., N'Diaye, E. N., Collart-Dutilleul, V., Maridonneau- Persson, H. L., Vainikka, L. K., Sege, M., Wennerstrom, U., Dam-Larsen, S., & Persson, J.
Parini, I., et al. (1999). The new ketolide HMR3647 accumulates in the azurophil (2012). Leaky lysosomes in lung transplant macrophages: azithromycin prevents ox-
granules of human polymorphonuclear cells. Antimicrob Agents Chemother 43, idative damage. Respir Res 13, 83.
2457–2462. Piacentini, G. L., Peroni, D. G., Bodini, A., Pigozzi, R., Costella, S., Loiacono, A., et al. (2007).
Miyazaki, M., Zaitsu, M., Honjo, K., Ishii, E., & Hamasaki, Y. (2003). Macrolide antibiotics Azithromycin reduces bronchial hyperresponsiveness and neutrophilic airway in-
inhibit prostaglandin E2 synthesis and mRNA expression of prostaglandin synthetic flammation in asthmatic children: a preliminary report. Allergy Asthma Proc 28,
enzymes in human leukocytes. Prostaglandins Leukot Essent Fatty Acids 69, 229–235. 194–198.
Mizunoe, S., Kadota, J., Tokimatsu, I., Kishi, K., Nagai, H., & Nasu, M. (2004). Clarithromycin Pinto, L. A., Pitrez, P. M., Luisi, F., de Mello, P. P., Gerhardt, M., Ferlini, R., et al. (2012).
and azithromycin induce apoptosis of activated lymphocytes via down-regulation of Azithromycin therapy in hospitalized infants with acute bronchiolitis is not associat-
Bcl-xL. Int Immunopharmacol 4, 1201–1207. ed with better clinical outcomes: a randomized, double-blinded, and placebo-
Montenez, J. P., Van Bambeke, F., Piret, J., Brasseur, R., Tulkens, P. M., & Mingeot-Leclercq, controlled clinical trial. J Pediatr 161, 1104–1108.
M. P. (1999). Interactions of macrolide antibiotics (Erythromycin A, roxithromycin, Pitsouni, E., Iavazzo, C., Athanasiou, S., & Falagas, M. E. (2007). Single-dose azithromycin
erythromycylamine [Dirithromycin], and azithromycin) with phospholipids: versus erythromycin or amoxicillin for Chlamydia trachomatis infection during preg-
computer-aided conformational analysis and studies on acellular and cell culture nancy: a meta-analysis of randomised controlled trials. Int J Antimicrob Agents 30,
models. Toxicol Appl Pharmacol 156, 129–140. 213–221.
Montenez, J. P., Van Bambeke, F., Piret, J., Schanck, A., Brasseur, R., Tulkens, P. M., et al. Plesko, S., Banic, M., Plecko, V., Anic, B., Brkic, T., Renata, H., et al. (2010). Effect of
(1996). Interaction of the macrolide azithromycin with phospholipids. II. Biophysical azithromycin on acute inflammatory lesions and colonic bacterial load in a murine
and computer-aided conformational studies. Eur J Pharmacol 314, 215–227. model of experimental colitis. Dig Dis Sci 55, 2211–2218.
Morinaga, Y., Yanagihara, K., Miyashita, N., Seki, M., Izumikawa, K., Kakeya, H., et al. Polancec, D. S., Munic Kos, V., Banjanac, M., Vrancic, M., Cuzic, S., Belamaric, D., et al.
(2009). Azithromycin, clarithromycin and telithromycin inhibit MUC5AC induction (2012). Azithromycin drives in vitro GM-CSF/IL-4-induced differentiation of human
by Chlamydophila pneumoniae in airway epithelial cells. Pulm Pharmacol Ther 22, blood monocytes toward dendritic-like cells with regulatory properties. J Leukoc
580–586. Biol 91, 229–243.
Moriya, S., Che, X. F., Komatsu, S., Abe, A., Kawaguchi, T., Gotoh, A., et al. (2013). Macrolide Pukander, J., & Rautianen, M. (1996). Penetration of azithromycin into middle ear effu-
antibiotics block autophagy flux and sensitize to bortezomib via endoplasmic reticu- sions in acute and secretory otitis media in children. J Antimicrob Chemother
lum stress-mediated CHOP induction in myeloma cells. Int J Oncol 42, 1541–1550. 37(Suppl. C), 53–61.
Mtairag, E. M., Abdelghaffar, H., Douhet, C., & Labro, M. T. (1995). Role of extracellular cal- Raines, D. A., Yusuf, A., Jabak, M. H., Ahmed, W. S., Karcioglu, Z. A., & El-Yazigi, A. (1998).
cium in in vitro uptake and intraphagocytic location of macrolides. Antimicrob Agents Simultaneous high-performance liquid chromatography analysis of azithromycin and
Chemother 39, 1676–1682. two of its metabolites in human tears and plasma. Ther Drug Monit 20, 680–684.
Mulholland, S., Gavranich, J. B., Gillies, M. B., & Chang, A. B. (2012). Antibiotics for Ramsey, P. S., Vaules, M. B., Vasdev, G. M., Andrews, W. W., & Ramin, K. D. (2003). Mater-
community-acquired lower respiratory tract infections secondary to Mycoplasma nal and transplacental pharmacokinetics of azithromycin. Am J Obstet Gynecol 188,
pneumoniae in children. Cochrane Database Syst Rev 9, CD004875. 714–718.
Munic, V., Banjanac, M., Kostrun, S., Nujic, K., Bosnar, M., Marjanovic, N., et al. (2011). In- Randow, F., & Munz, C. (2012). Autophagy in the regulation of pathogen replication and
tensity of macrolide anti-inflammatory activity in J774A.1 cells positively correlates adaptive immunity. Trends Immunol 33, 475–487.
with cellular accumulation and phospholipidosis. Pharmacol Res 64, 298–307. Ray, W. A., Murray, K. T., Hall, K., Arbogast, P. G., & Stein, C. M. (2012). Azithromycin and
Munic, V., Kelneric, Z., Mikac, L., & Erakovic Haber, V. (2010). Differences in assessment of the risk of cardiovascular death. N Engl J Med 366, 1881–1890.
macrolide interaction with human MDR1 (ABCB1, P-gp) using rhodamine-123 efflux, Renna, M., Schaffner, C., Brown, K., Shang, S., Tamayo, M. H., Hegyi, K., et al. (2011).
ATPase activity and cellular accumulation assays. Eur J Pharm Sci 41, 86–95. Azithromycin blocks autophagy and may predispose cystic fibrosis patients to myco-
Murphy, B. S., Sundareshan, V., Cory, T. J., Hayes, D., Jr., Anstead, M. I., & Feola, D. J. (2008). bacterial infection. J Clin Invest 121, 3554–3563.
Azithromycin alters macrophage phenotype. J Antimicrob Chemother 61, 554–560. Restrepo, M. I., Mortensen, E. M., Waterer, G. W., Wunderink, R. G., Coalson, J. J., &
Murphy, D. M., Forrest, I. A., Corris, P. A., Johnson, G. E., Small, T., Jones, D., et al. (2008). Anzueto, A. (2009). Impact of macrolide therapy on mortality for patients with severe
Azithromycin attenuates effects of lipopolysaccharide on lung allograft bronchial ep- sepsis due to pneumonia. Eur Respir J 33, 153–159.
ithelial cells. J Heart Lung Transplant 27, 1210–1216. Ribeiro, C. M., Hurd, H., Wu, Y., Martino, M. E., Jones, L., Brighton, B., et al. (2009).
Murray, J. J., Emparanza, P., Lesinskas, E., Tawadrous, M., & Breen, J. D. (2005). Efficacy and Azithromycin treatment alters gene expression in inflammatory, lipid metabolism,
safety of a novel, single-dose azithromycin microsphere formulation versus 10 days and cell cycle pathways in well-differentiated human airway epithelia. PLoS One 4,
of levofloxacin for the treatment of acute bacterial sinusitis in adults. Otolaryngol e5806.
Head Neck Surg 133, 194–200. Rodvold, K. A., Danziger, L. H., & Gotfried, M. H. (2003). Steady-state plasma and
Muto, C., Liu, P., Chiba, K., & Suwa, T. (2011). Pharmacokinetic–pharmacodynamic analy- bronchopulmonary concentrations of intravenous levofloxacin and azithromycin in
sis of azithromycin extended release in Japanese patients with common respiratory healthy adults. Antimicrob Agents Chemother 47, 2450–2457.
tract infectious disease. J Antimicrob Chemother 66, 165–174. Rohof, W. O., Bennink, R. J., de Ruigh, A. A., Hirsch, D. P., Zwinderman, A. H., &
Naidoo, R. V., & Bryant, P. A. (2009). Not every cough in bronchiolitis season is bronchiol- Boeckxstaens, G. E. (2012). Effect of azithromycin on acid reflux, hiatus hernia and
itis. BMJ Case Rep 2009. proximal acid pocket in the postprandial period. Gut 61, 1670–1677.
Navarro-Xavier, R. A., Newson, J., Silveira, V. L., Farrow, S. N., Gilroy, D. W., & Bystrom, J. Saiman, L., Anstead, M., Mayer-Hamblett, N., Lands, L. C., Kloster, M., Hocevar-Trnka, J.,
(2010). A new strategy for the identification of novel molecules with targeted et al. (2010). Effect of azithromycin on pulmonary function in patients with cystic fi-
proresolution of inflammation properties. J Immunol 184, 1516–1525. brosis uninfected with Pseudomonas aeruginosa: a randomized controlled trial. JAMA
Nie, Y. C., Wu, H., Li, P. B., Xie, L. M., Luo, Y. L., Shen, J. G., et al. (2012). Naringin attenuates 303, 1707–1715.
EGF-induced MUC5AC secretion in A549 cells by suppressing the cooperative activi- Saiman, L., Marshall, B. C., Mayer-Hamblett, N., Burns, J. L., Quittner, A. L., Cibene, D. A.,
ties of MAPKs-AP-1 and IKKs-IkappaB-NF-kappaB signaling pathways. Eur J et al. (2003). Azithromycin in patients with cystic fibrosis chronically infected with
Pharmacol 690, 207–213. Pseudomonas aeruginosa: a randomized controlled trial. JAMA 290, 1749–1756.
Nightingale, C. H. (1997). Pharmacokinetics and pharmacodynamics of newer macrolides. Saint-Criq, V., Ruffin, M., Rebeyrol, C., Guillot, L., Jacquot, J., Clement, A., et al. (2012).
Pediatr Infect Dis J 16, 438–443. Azithromycin fails to reduce inflammation in cystic fibrosis airway epithelial cells.
Nujic, K., Banjanac, M., Munic, V., Polancec, D., & Erakovic Haber, V. (2012). Impairment of Eur J Pharmacol 674, 1–6.
lysosomal functions by azithromycin and chloroquine contributes to anti- Sakito, O., Kadota, J., Kohno, S., Abe, K., Shirai, R., & Hara, K. (1996). Interleukin 1 beta,
inflammatory phenotype. Cell Immunol 279, 78–86. tumor necrosis factor alpha, and interleukin 8 in bronchoalveolar lavage fluid of pa-
Nujic, K., Smith, M., Lee, M., Belamaric, D., Tomaskovic, L., Alihodzic, S., et al. (2012). tients with diffuse panbronchiolitis: a potential mechanism of macrolide therapy.
Valosin containing protein (VCP) interacts with macrolide antibiotics without medi- Respiration 63, 42–48.
ating their anti-inflammatory activities. Eur J Pharmacol 677, 163–172. Salman, S., Rogerson, S. J., Kose, K., Griffin, S., Gomorai, S., Baiwog, F., et al. (2010). Phar-
Panpanich, R., Lerttrakarnnon, P., & Laopaiboon, M. (2008). Azithromycin for acute lower macokinetic properties of azithromycin in pregnancy. Antimicrob Agents Chemother
respiratory tract infections. Cochrane Database Syst Rev, CD001954. 54, 360–366.

Please cite this article as: Parnham, M.J., et al., Azithromycin: Mechanisms of action and their relevance for clinical applications, Pharmacology &
Therapeutics (2014), http://dx.doi.org/10.1016/j.pharmthera.2014.03.003
20 M.J. Parnham et al. / Pharmacology & Therapeutics xxx (2014) xxx–xxx

Sato, Y., Kaneko, K., & Inoue, M. (2007). Macrolide antibiotics promote the LPS-induced pneumococcal strains selected in vitro by macrolide passage. Antimicrob Agents
upregulation of prostaglandin E receptor EP2 and thus attenuate macrolide suppres- Chemother 44, 2118–2125.
sion of IL-6 production. Prostaglandins Leukot Essent Fatty Acids 76, 181–188. Tamaoki, J., Kadota, J., & Takizawa, H. (2004). Clinical implications of the immunomodu-
Scaglione, F., Demartini, G., Dugnani, S., Arcidiacono, M. M., Pintucci, J. P., & Fraschini, F. latory effects of macrolides. Am J Med 117(Suppl. 9A), 5S–11S.
(1999). Interpretation of middle ear fluid concentrations of antibiotics: comparison Tateda, K., Comte, R., Pechere, J. C., Kohler, T., Yamaguchi, K., & Van Delden, C. (2001).
between ceftibuten, cefixime and azithromycin. Br J Clin Pharmacol 47, 267–271. Azithromycin inhibits quorum sensing in Pseudomonas aeruginosa. Antimicrob
Serisier, D. J., & Martin, M. L. (2011). Long-term, low-dose erythromycin in bronchiectasis Agents Chemother 45, 1930–1933.
subjects with frequent infective exacerbations. Respir Med 105, 946–949. Terao, H., Asano, K., Kanai, K., Kyo, Y., Watanabe, S., Hisamitsu, T., et al. (2003). Suppres-
Shakeri-Nejad, K., & Stahlmann, R. (2006). Drug interactions during therapy with three sive activity of macrolide antibiotics on nitric oxide production by lipopolysaccharide
major groups of antimicrobial agents. Expert Opin Pharmacother 7, 639–651. stimulation in mice. Mediators Inflamm 12, 195–202.
Shepard, R. M., & Falkner, F. C. (1990). Pharmacokinetics of azithromycin in rats and dogs. Tessmer, A., Welte, T., Martus, P., Schnoor, M., Marre, R., & Suttorp, N. (2009). Impact of
J Antimicrob Chemother 25(Suppl. A), 49–60. intravenous {beta}-lactam/macrolide versus {beta}-lactam monotherapy on mortali-
Shinkai, M., Henke, M. O., & Rubin, B. K. (2008). Macrolide antibiotics as immunomodula- ty in hospitalized patients with community-acquired pneumonia. J Antimicrob
tory medications: proposed mechanisms of action. Pharmacol Ther 117, 393–405. Chemother 63, 1025–1033.
Shinkai, M., Tamaoki, J., Kobayashi, H., Kanoh, S., Motoyoshi, K., Kute, T., et al. (2006). Thorpe, E. M., Jr., Stamm, W. E., Hook, E. W., III, Gall, S. A., Jones, R. B., Henry, K., et al.
Clarithromycin delays progression of bronchial epithelial cells from G1 phase to S (1996). Chlamydial cervicitis and urethritis: single dose treatment compared with
phase and delays cell growth via extracellular signal-regulated protein kinase sup- doxycycline for seven days in community based practises. Genitourin Med 72, 93–97.
pression. Antimicrob Agents Chemother 50, 1738–1744. Tomaskovic, L., Komac, M., Makaruha Stegic, O., Munic, V., Ralic, J., Stanic, B., et al. (2013).
Shorr, A. F., Zilberberg, M. D., Kan, J., Hoffman, J., Micek, S. T., & Kollef, M. H. (2013). Macrolactonolides: a novel class of anti-inflammatory compounds. Bioorg Med Chem
Azithromycin and survival in Streptococcus pneumoniae pneumonia: a retrospective 21, 321–332.
study. BMJ Open 3. Tomazic, J., Kotnik, V., & Wraber, B. (1993). In vivo administration of azithromycin affects
Singlas, E. (1995). Clinical pharmacokinetics of azithromycin. Pathol Biol (Paris) 43, lymphocyte activity in vitro. Antimicrob Agents Chemother 37, 1786–1789.
505–511. Tong, J., Liu, Z. C., & Wang, D. X. (2011). Azithromycin acts as an immunomodulatory
Skalsky, K., Yahav, D., Lador, A., Eliakim-Raz, N., Leibovici, L., & Paul, M. (2013). Macrolides agent to suppress the expression of TREM-1 in Bacillus pyocyaneus-induced sepsis.
vs. quinolones for community-acquired pneumonia: meta-analysis of randomized Immunol Lett 138, 137–143.
controlled trials. Clin Microbiol Infect 19, 370–378. Tresse, E., Salomons, F. A., Vesa, J., Bott, L. C., Kimonis, V., Yao, T. P., et al. (2010). VCP/p97 is
Skerk, V., Krhen, I., Lisic, M., Begovac, J., Roglic, S., Skerk, V., et al. (2004). Comparative ran- essential for maturation of ubiquitin-containing autophagosomes and this function is
domized pilot study of azithromycin and doxycycline efficacy in the treatment of impaired by mutations that cause IBMPFD. Autophagy 6, 217–227.
prostate infection caused by Chlamydia trachomatis. Int J Antimicrob Agents 24, Tsai, W. C., Hershenson, M. B., Zhou, Y., & Sajjan, U. (2009). Azithromycin increases surviv-
188–191. al and reduces lung inflammation in cystic fibrosis mice. Inflamm Res 58, 491–501.
Skerk, V., Marekovic, I., Markovinovic, L., Begovac, J., Skerk, V., Barsic, N., et al. (2006). Tsai, W. C., Rodriguez, M. L., Young, K. S., Deng, J. C., Thannickal, V. J., Tateda, K., et al.
Comparative randomized pilot study of azithromycin and doxycycline efficacy and (2004). Azithromycin blocks neutrophil recruitment in Pseudomonas endobronchial
tolerability in the treatment of prostate infection caused by Ureaplasma urealyticum. infection. Am J Respir Crit Care Med 170, 1331–1339.
Chemotherapy 52, 9–11. Tsai, W. C., & Standiford, T. J. (2004). Immunomodulatory effects of macrolides in the
Skerk, V., Schonwald, S., Krhen, I., Banaszak, A., Begovac, J., Strugar, J., et al. (2003). Com- lung: lessons from in-vitro and in-vivo models. Curr Pharm Des 10, 3081–3093.
parative analysis of azithromycin and ciprofloxacin in the treatment of chronic pros- Tyteca, D., Schanck, A., Dufrene, Y. F., Deleu, M., Courtoy, P. J., Tulkens, P. M., et al. (2003).
tatitis caused by Chlamydia trachomatis. Int J Antimicrob Agents 21, 457–462. The macrolide antibiotic azithromycin interacts with lipids and affects membrane or-
Skerk, V., Schonwald, S., Krhen, I., Markovinovic, L., Barsic, B., Marekovic, I., et al. (2002). ganization and fluidity: studies on Langmuir–Blodgett monolayers, liposomes and
Comparative analysis of azithromycin and clarithromycin efficacy and tolerability in J774 macrophages. J Membr Biol 192, 203–215.
the treatment of chronic prostatitis caused by Chlamydia trachomatis. J Chemother Tyteca, D., Van Der Smissen, P., Mettlen, M., Van Bambeke, F., Tulkens, P. M., Mingeot-
14, 384–389. Leclercq, M. P., et al. (2002). Azithromycin, a lysosomotropic antibiotic, has distinct
Southern, K. W., Barker, P. M., Solis-Moya, A., & Patel, L. (2011). Macrolide antibiotics for effects on fluid-phase and receptor-mediated endocytosis, but does not impair
cystic fibrosis. Cochrane Database Syst Rev, CD002203. phagocytosis in J774 macrophages. Exp Cell Res 281, 86–100.
Spagnolo, P., Fabbri, L. M., & Bush, A. (2013). Long-term macrolide treatment for chronic Tyteca, D., Van Der Smissen, P., Van Bambeke, F., Leys, K., Tulkens, P. M., Courtoy, P. J., et al.
respiratory disease. Eur Respir J 42, 239–251. (2001). Azithromycin, a lysosomotropic antibiotic, impairs fluid-phase pinocytosis in
Spyridaki, A., Raftogiannis, M., Antonopoulou, A., Tsaganos, T., Routsi, C., Baziaka, F., et al. cultured fibroblasts. Eur J Cell Biol 80, 466–478.
(2012). Effect of clarithromycin in inflammatory markers of patients with ventilator- Valery, P. C., Morris, P. S., Byrnes, C. A., Grimwood, K., Torzillo, P. J., Bauert, P. A., et al. (2013).
associated pneumonia and sepsis caused by Gram-negative bacteria: results from a Long-term azithromycin for Indigenous children with non-cystic-fibrosis bronchiecta-
randomized clinical study. Antimicrob Agents Chemother 56, 3819–3825. sis or chronic suppurative lung disease (Bronchiectasis Intervention Study): a
Srivastava, P., Vardhan, H., Bhengraj, A. R., Jha, R., Singh, L. C., Salhan, S., et al. (2011). multicentre, double-blind, randomised controlled trial. Lancet Respir Med 1, 610–620.
Azithromycin treatment modulates the extracellular signal-regulated kinase mediat- Van Bambeke, F., Montenez, J. P., Piret, J., Tulkens, P. M., Courtoy, P. J., & Mingeot-Leclercq,
ed pathway and inhibits inflammatory cytokines and chemokines in epithelial cells M. P. (1996). Interaction of the macrolide azithromycin with phospholipids. I. Inhibi-
from infertile women with recurrent Chlamydia trachomatis infection. DNA Cell Biol tion of lysosomal phospholipase A1 activity. Eur J Pharmacol 314, 203–214.
30, 545–554. van Delden, C., Kohler, T., Brunner-Ferber, F., Francois, B., Carlet, J., & Pechere, J. C. (2012).
Stamatiou, R., Boukas, K., Paraskeva, E., Molyvdas, P. A., & Hatziefthimiou, A. (2010). Azithromycin to prevent Pseudomonas aeruginosa ventilator-associated pneumonia
Azithromycin reduces the viability of human bronchial smooth muscle cells. J by inhibition of quorum sensing: a randomized controlled trial. Intensive Care Med
Antibiot (Tokyo) 63, 71–75. 38, 1118–1125.
Stamatiou, R., Paraskeva, E., Boukas, K., Gourgoulianis, K. I., Molyvdas, P. A., & van Eijk, A. M., & Terlouw, D. J. (2011). Azithromycin for treating uncomplicated malaria.
Hatziefthimiou, A. A. (2009). Azithromycin has an antiproliferative and autophagic Cochrane Database Syst Rev, CD006688.
effect on airway smooth muscle cells. Eur Respir J 34, 721–730. Vanaudenaerde, B. M., Meyts, I., Vos, R., Geudens, N., De Wever, W., Verbeken, E. K., et al.
Stamm, W. E., Hicks, C. B., Martin, D. H., Leone, P., Hook, E. W., III, Cooper, R. H., et al. (2008). A dichotomy in bronchiolitis obliterans syndrome after lung transplantation
(1995). Azithromycin for empirical treatment of the nongonococcal urethritis syn- revealed by azithromycin therapy. Eur Respir J 32, 832–843.
drome in men. A randomized double-blind study. JAMA 274, 545–549. Vanaudenaerde, B. M., Vos, R., Meyts, I., De Vleeschauwer, S. I., Verleden, S. E., Widyastuti-
Stepanic, V., Kostrun, S., Malnar, I., Hlevnjak, M., Butkovic, K., Caleta, I., et al. (2011). Willems, A., et al. (2008). Macrolide therapy targets a specific phenotype in respira-
Modeling cellular pharmacokinetics of 14- and 15-membered macrolides with phys- tory medicine: from clinical experience to basic science and back. Inflamm Allergy
icochemical properties. J Med Chem 54, 719–733. Drug Targets 7, 279–287.
Stover, D. E., & Mangino, D. (2005). Macrolides: a treatment alternative for bronchiolitis Vanaudenaerde, B. M., Wuyts, W. A., Geudens, N., Dupont, L. J., Schoofs, K., Smeets, S., et al.
obliterans organizing pneumonia? Chest 128, 3611–3617. (2007). Macrolides inhibit IL17-induced IL8 and 8-isoprostane release from human
Sugie, M., Asakura, E., Zhao, Y. L., Torita, S., Nadai, M., Baba, K., et al. (2004). Possible in- airway smooth muscle cells. Am J Transplant 7, 76–82.
volvement of the drug transporters P glycoprotein and multidrug resistance- Verleden, G. M., Vanaudenaerde, B. M., Dupont, L. J., & Van Raemdonck, D. E. (2006).
associated protein Mrp2 in disposition of azithromycin. Antimicrob Agents Azithromycin reduces airway neutrophilia and interleukin-8 in patients with bron-
Chemother 48, 809–814. chiolitis obliterans syndrome. Am J Respir Crit Care Med 174, 566–570.
Sugiyama, K., Shirai, R., Mukae, H., Ishimoto, H., Nagata, T., Sakamoto, N., et al. (2007). Dif- Verleden, S. E., Vandooren, J., Vos, R., Willems, S., Dupont, L. J., Verleden, G. M., et al.
fering effects of clarithromycin and azithromycin on cytokine production by murine (2011). Azithromycin decreases MMP-9 expression in the airways of lung transplant
dendritic cells. Clin Exp Immunol 147, 540–546. recipients. Transpl Immunol 25, 159–162.
Suresh Babu, K., Kastelik, J., & Morjaria, J. B. (2013). Role of long term antibiotics in chronic Verleden, S. E., Vos, R., Mertens, V., Willems-Widyastuti, A., De Vleeschauwer, S. I., Dupont, L.
respiratory diseases. Respir Med 107, 800–815. J., et al. (2011). Heterogeneity of chronic lung allograft dysfunction: insights from pro-
Sutcliffe, J. A., & Leclerq, R. (2002). Mechanisms of resistance to macrolides, lincosamides, tein expression in broncho alveolar lavage. J Heart Lung Transplant 30, 667–673.
and ketolides. In W. Schönfeld, & H. A. Kirst (Eds.), Macrolide Antibiotics Videler, W. J., Badia, L., Harvey, R. J., Gane, S., Georgalas, C., van der Meulen, F. W., et al.
(pp. 281–317). Basel: Birkhäuser Verlag. (2011). Lack of efficacy of long-term, low-dose azithromycin in chronic
Svanstrom, H., Pasternak, B., & Hviid, A. (2013). Use of azithromycin and death from car- rhinosinusitis: a randomized controlled trial. Allergy 66, 1457–1468.
diovascular causes. N Engl J Med 368, 1704–1712. Villarino, N., & Martin-Jimenez, T. (2013). Pharmacokinetics of macrolides in foals. J Vet
Tabbara, K. F., al-Kharashi, S. A., al-Mansouri, S. M., al-Omar, O. M., Cooper, H., el-Asrar, A. Pharmacol Ther 36, 1–13.
M., et al. (1998). Ocular levels of azithromycin 116, 1625–1628. Vos, R., Vanaudenaerde, B. M., Ottevaere, A., Verleden, S. E., De Vleeschauwer, S. I.,
Tait-Kamradt, A., Davies, T., Cronan, M., Jacobs, M. R., Appelbaum, P. C., & Sutcliffe, J. Willems-Widyastuti, A., et al. (2010). Long-term azithromycin therapy for bronchiol-
(2000). Mutations in 23S rRNA and ribosomal protein L4 account for resistance in itis obliterans syndrome: divide and conquer? J Heart Lung Transplant 29, 1358–1368.

Please cite this article as: Parnham, M.J., et al., Azithromycin: Mechanisms of action and their relevance for clinical applications, Pharmacology &
Therapeutics (2014), http://dx.doi.org/10.1016/j.pharmthera.2014.03.003
M.J. Parnham et al. / Pharmacology & Therapeutics xxx (2014) xxx–xxx 21

Vos, R., Vanaudenaerde, B. M., Verleden, S. E., De Vleeschauwer, S. I., Willems-Widyastuti, Wong, C., Jayaram, L., Karalus, N., Eaton, T., Tong, C., Hockey, H., et al. (2012).
A., Van Raemdonck, D. E., et al. (2011). A randomised controlled trial of azithromycin Azithromycin for prevention of exacerbations in non-cystic fibrosis bronchiectasis
to prevent chronic rejection after lung transplantation. Eur Respir J 37, 164–172. (EMBRACE): a randomised, double-blind, placebo-controlled trial. Lancet 380,
Vrancic, M., Banjanac, M., Nujic, K., Bosnar, M., Murati, T., Munic, V., et al. (2012). 660–667.
Azithromycin distinctively modulates classical activation of human monocytes Wuyts, W. A., Willems, S., Vos, R., Vanaudenaerde, B. M., De Vleeschauwer, S. I., Rinaldi, M.,
in vitro. Br J Pharmacol 165, 1348–1360. et al. (2010). Azithromycin reduces pulmonary fibrosis in a bleomycin mouse model.
Wang, L., Kitaichi, K., Hui, C. S., Takagi, K., Takagi, K., Sakai, M., et al. (2000). Reversal of Exp Lung Res 36, 602–614.
anticancer drug resistance by macrolide antibiotics in vitro and in vivo. Clin Exp Yamada, K., Yanagihara, K., Kaku, N., Harada, Y., Migiyama, Y., Nagaoka, K., et al. (2013).
Pharmacol Physiol 27, 587–593. Azithromycin attenuates lung inflammation in a mouse model of ventilator-
Wang, P. L. (2010). Roles of oral bacteria in cardiovascular diseases—from molecular associated pneumonia by multidrug-resistant Acinetobacter baumannii. Antimicrob
mechanisms to clinical cases: treatment of periodontal disease regarded as biofilm Agents Chemother 57, 3883–3888.
infection: systemic administration of azithromycin. J Pharmacol Sci 113, 126–133. Yamamoto, H., Naito, Y., Okano, M., Kanazawa, T., Takematsu, H., & Kozutsumi, Y.
Wierzbowski, A. K., Hoban, D. J., Hisanaga, T., DeCorby, M., & Zhanel, G. G. (2006). The use (2011). Sphingosylphosphorylcholine and lysosulfatide have inverse regulatory
of macrolides in treatment of upper respiratory tract infections. Curr Allergy Asthma functions in monocytic cell differentiation into macrophages. Arch Biochem
Rep 6, 171–181. Biophys 506, 83–91.
Wildfeuer, A., Reisert, I., & Laufen, H. (1993). Uptake and subcellular distribution of Yamauchi, K., Shibata, Y., Kimura, T., Abe, S., Inoue, S., Osaka, D., et al. (2009).
azithromycin in human phagocytic cells. Demonstration of the antibiotic in neutro- Azithromycin suppresses interleukin-12p40 expression in lipopolysaccharide and
phil polymorphonuclear leucocytes and monocytes by autoradiography and electron interferon-gamma stimulated macrophages. Int J Biol Sci 5, 667–678.
microscopy. Arzneimittelforschung 43, 484–486. Yamaya, M., Azuma, A., Takizawa, H., Kadota, J., Tamaoki, J., & Kudoh, S. (2012).
Willems-Widyastuti, A., Vanaudenaerde, B. M., Vos, R., Dilisen, E., Verleden, S. E., De Macrolide effects on the prevention of COPD exacerbations. Eur Respir J 40,
Vleeschauwer, S. I., et al. (2013). Azithromycin attenuates fibroblast growth factors 485–494.
induced vascular endothelial growth factor via p38(MAPK) signaling in human air- Yanagihara, K., Izumikawa, K., Higa, F., Tateyama, M., Tokimatsu, I., Hiramatsu, K., et al.
way smooth muscle cells. Cell Biochem Biophys 67, 331–339. (2009). Efficacy of azithromycin in the treatment of community-acquired pneumo-
Willenborg, M., Schmidt, C. K., Braun, P., Landgrebe, J., von Figura, K., Saftig, P., et al. nia, including patients with macrolide-resistant Streptococcus pneumoniae infection.
(2005). Mannose 6-phosphate receptors, Niemann-Pick C2 protein, and lysosomal Intern Med 48, 527–535.
cholesterol accumulation. J Lipid Res 46, 2559–2569. Yang, M., Dong, B. R., Lu, J., Lin, X., & Wu, H. M. (2013). Macrolides for diffuse
Wilms, E. B., Touw, D. J., & Heijerman, H. G. (2006). Pharmacokinetics of azithromycin in panbronchiolitis. Cochrane Database Syst Rev 2, CD007716.
plasma, blood, polymorphonuclear neutrophils and sputum during long-term thera- Yasuda, K., Lan, L. B., Sanglard, D., Furuya, K., Schuetz, J. D., & Schuetz, E. G. (2002). Inter-
py in patients with cystic fibrosis. Ther Drug Monit 28, 219–225. action of cytochrome P450 3A inhibitors with P-glycoprotein. J Pharmacol Exp Ther
Wilson, B. Z., Anzueto, A., Restrepo, M. I., Pugh, M. J., & Mortensen, E. M. (2012). Compar- 303, 323–332.
ison of two guideline-concordant antimicrobial combinations in elderly patients hos- Yeung, T., & Grinstein, S. (2007). Lipid signaling and the modulation of surface charge
pitalized with severe community-acquired pneumonia. Crit Care Med 40, 2310–2314. during phagocytosis. Immunol Rev 219, 17–36.

Please cite this article as: Parnham, M.J., et al., Azithromycin: Mechanisms of action and their relevance for clinical applications, Pharmacology &
Therapeutics (2014), http://dx.doi.org/10.1016/j.pharmthera.2014.03.003

Vous aimerez peut-être aussi