You are on page 1of 20

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/261881044

Promises and failures of gallium as an antibacterial agent

Article  in  Future Microbiology · March 2014


DOI: 10.2217/fmb.14.3 · Source: PubMed

CITATIONS READS

54 1,446

5 authors, including:

Fabrizia Minandri Emanuela Frangipani


Università Degli Studi Roma Tre Università degli Studi di Urbino "Carlo Bo"
19 PUBLICATIONS   436 CITATIONS    56 PUBLICATIONS   792 CITATIONS   

SEE PROFILE SEE PROFILE

Francesco Imperi Paolo Visca


Università Degli Studi Roma Tre Università Degli Studi Roma Tre
81 PUBLICATIONS   2,613 CITATIONS    402 PUBLICATIONS   11,577 CITATIONS   

SEE PROFILE SEE PROFILE

Some of the authors of this publication are also working on these related projects:

Development of a POCT medical device for fast diagnosis and improved management of urinary tract infections (UTIs) View project

Microbial persistence View project

All content following this page was uploaded by Francesco Imperi on 30 April 2014.

The user has requested enhancement of the downloaded file.


Review

014
Promises and failures of gallium as an
antibacterial agent

Fabrizia Minandri1, Carlo Bonchi1, Emanuela Frangipani1, Francesco Imperi2


& Paolo Visca*,1

of
Abstract: Gallium has a long history as a diagnostic and chemotherapeutic agent. The
pharmacological properties of Ga(III) rely on chemical mimicry; when Ga(III) is exogenously
supplied to living cells it can replace Fe(III) within target molecules, thereby perturbing
bacterial metabolism. Ga(III)-induced metabolic distresses are dramatic in fast-growing

ro
& Visca

cells, like bacterial cells. Interest in the antibacterial properties of Ga(III) has been raised by
the compelling need for novel drugs to combat multidrug-resistant bacteria and by the
shortage of new antibiotic candidates in the pharmaceutical pipeline. Ga(III) activity has
been demonstrated, both in vitro and in animal models of infections, on several bacterial
rP
pathogens, also including intracellular and biofilm-forming bacteria. Ga(III) activity is
affected by iron availability and the metabolic state of the cell, being maximal in iron-poor
media and in respiring cells. Synergism between Ga(III) and antibiotics holds promise as last
B.14.3
resort therapy for infections sustained by pandrug-resistant bacteria.
ho

uture Background Keywords 


Metals such as copper, mercury and silver have been known for their ability to suppress microbial • Acinetobacter • antibiotic
growth since antiquity. The antimicrobial properties of metals have been theorized by the Swiss chemist resistance • anti-infectives
lium as Nägeli [1] with the ‘oligodynamic effect’ (from the Greek: oligos = few, dynamis = force), which is a • Burkholderia • Ganite®
toxic effect observed at very low concentrations. Metal compounds have been exploited for a number of • iron metabolism
clinical applications, and have become an essential component of the anti-infective armamentarium [2] . • Mycobacterium
ut

The semimetal gallium (Ga) is not novel to clinical applications, as it has been extensively used as a • Pseudomonas
diagnostic for the localization of malignant cells and inflammatory lesions, or as a therapeutic for the • Staphylococcus
treatment of cancer, autoimmune diseases and bone resorption disorders [3–7] . The pharmacological
properties of Ga(III) are attributable to its similarity with Fe(III) at the level of nuclear radius, coordi-
A

nation chemistry and ionization potential. These features allow Ga(III) to substitute for Fe(III) in the
prosthetic group of several enzymes (reviewed in [8]). Iron transport proteins, such as transferrin (Tf)
and lactoferrin (Lf), are able to form complexes with Ga(III) and deliver it to the cell [9,10] . Inside the
cell, Ga(III) perturbs iron metabolism since it is incorporated into essential proteins and enzymes in
place of iron. Unlike iron, Ga(III) cannot be reduced under physiological conditions and, therefore,
it cannot take part in redox reactions, ultimately inhibiting a number of essential functions. Several
metabolic pathways, involving both Fe(III)-dependent and -independent enzymes, are targeted by
Ga(III), and Ga(III)-induced metabolic distresses are dramatic in fast-growing cells, like cancer and
bacterial cells [3,11–17] .
y Ga(III), in the form of citrate-buffered gallium nitrate [Ga(NO3)3], is approved by the US FDA
(Ganite®, Genta, NJ, USA) for the treatment of cancer-associated hypercalcemia. Other gallium
ure
UK 1
Department of Sciences, University “Roma Tre”, Viale Guglielmo Marconi 446, 00146 Rome, Italy,
2
Department of Biology & Biotechnology “C. Darwin”, “Sapienza” University of Rome, Rome, Italy
*Author for correspondence: Tel.: +39 06 57336347; Fax: +39 06 57336321; paolo.visca@uniroma3.it part of

10.2217/FMB.14.3 © 2014 Future Medicine Ltd Future Microbiol. (2014) 9(3), 379–397 ISSN 1746-0913 379
Review  Minandri, Bonchi, Frangipani, Imperi & Visca

compounds, such as gallium chloride and gal- Under anaerobic or reducing conditions,
lium maltolate, have been developed to improve Fe(II) is the prevalent iron form and, different
antimetabolite activity while mitigating toxic from Fe(III), it freely diffuses through outer
effects, and some of these compounds are under membrane porins of Gram-negative bacteria,
preclinical investigation [18,19] . and it is transported through the cytoplasmic
The importance of iron in bacterial metabo- membrane by an ABC Fe(II) transporter, called
lism and virulence has fostered the exploitation Feo (Figure 1) [38] .
of Ga(III) as an anti-infective [20–30] . Hereafter, Iron acquisition must be finely controlled,
we will provide a comprehensive review of recent since both iron deficiency and iron overload
findings on the in vitro and in vivo antibacterial can be detrimental to cells. In fact, while too
activity of Ga(III), and discuss the pros and cons low intracellular iron cannot support bacterial
of Ga(III) as an anti-infective. growth, too high intracellular Fe(II) concentra-
tions cause the production of hydroxyl radicals
Effect of gallium on bacterial metabolism (via Fenton-type reactions) and oxidative dam-
Ga(III) acts as an iron mimetic and is supposed age [31] . Therefore, bacteria must sense intracel-

of
to exploit Fe(III) transport systems to enter bac- lular iron levels, and regulate iron uptake accord-
terial cells. Basic information on Fe(III) uptake ingly. In all bacteria iron homeostasis is ensured
and metabolism in bacteria are therefore instru- by transcriptional repressors of the Fur or DtxR
mental to understanding the mechanism(s) of family, which require Fe(II) as a cofactor to

ro
action and uptake routes of Ga(III). efficiently bind the promoter of target genes [39]
(Figure 1) . In addition, small regulatory RNAs
●●Bacterial iron uptake & metabolism (sRNAs) controlling the expression of iron-con-
Iron is scarcely bioavailable under aerobic con- taining proteins in response to intracellular iron
ditions since it is almost completely present as level variations, have been identified in different
rP
an oxidized and insoluble oxyhydroxide Fe(III) bacterial species (Figure 1) (reviewed in [40,41]).
polymer [31] . Bacteria have evolved a multiplicity
of strategies to cope with low iron bioavailability. ●●Mechanism of action & uptake routes of
Almost all mechanisms rely on proteins at the cell gallium in bacteria
surface that bind and transport Fe(III) or heme- Given the similarities between Ga(III) and
ho

containing molecules, or Fe(II) under anaerobic Fe(III) ions and the inability of the former to
conditions. The many iron uptake strategies used undergo redox cycles, it is plausible that Ga(III)
by Gram-positive and Gram-negative bacteria can substitute for Fe(III) in iron-containing
have been exensively reviewed [32,33] and are sum- enzymes, thus repressing their activity [8] . This
marized in Figure 1. The Fe(III) and heme acqui- is consistent with the observation that, in any
sition systems of bacteria can either involve direct bacterial species investigated to date, the anti-
ut

contact between cells and Fe(III)/heme sources, microbial activity of Ga(III) is counteracted by
or the synthesis and secretion of siderophores and an excess of Fe(III). Since many iron-containing
hemophores, which scavenge Fe(III) and heme, enzymes are involved in critical functions in bac-
respectively, from natural sources and/or host teria, such as DNA synthesis and repair, metabo-
A

proteins [33] . Hemophore-based heme uptake is lism, respiration and oxidative stress response
exploited by several Gram-negative bacteria [34] [42] , Ga(III) is likely to cause multiple deleteri-
and by Bacillus anthracis [35] . Heme can either ous effects to bacterial cells. Whether the anti-
be incorporated into heme-requiring proteins bacterial activity of Ga(III) relies on a general
or cleaved by cytoplasmic heme oxygenases to perturbation of bacterial iron metabolism or on
release iron [36] . Conversely, Fe(III) must always the inhibition of a specific enzyme and/or cellu-
be dissociated from siderophores either by hydrol- lar pathway remains an open question. A recent
ysis, or modification of the siderophore scaffold, study on the antibacterial activity of Ga(III)
or reduction of Fe(III) to Fe(II). The majority of against Pseudomonas aeruginosa reported a strong
Fe(III)-siderophore complexes are internalized influence of the carbon source, indicating that
into the cytoplasm prior to iron release, although some metabolic pathways may be more iron-
some undergo periplasmic reduction of Fe(III) to demanding, and thus more Ga(III)-sensitive,
Fe(II), followed by Fe(II) transport into the cyto- than others [29] . Moreover, pioneering studies
sol by dedicated inner membrane transporters on the neighbour species Pseudomonas fluorescens
(reviewed in (Figure 1) [37]). revealed that bacterial cells adapted to grow in

380 Future Microbiol. (2014) 9(3) future science group


Promises & failures of gallium as an antibacterial agent  Review

Diderm bacteria
(Gram-negative)

TonB

TonB
ExbD
ExbB

ExbD
B

ExbB
n
To

bB
bD

Ex Iron-uptake
Ex

OM genes
PS
CM ryhB gene

CM
RyhB-like sRNAs

of
mRNAs for iron-using proteins

Monoderm bacteria
(Gram-postive)

Fe(III) Fe(II) Siderophore ro


Ferri-siderophore Heme Host iron-binding proteins
rP
Figure 1. Iron-uptake strategies and iron-dependent gene regulation in bacteria. Fe(III) is delivered to bacterial cells by host iron-
binding proteins (transferrin or lactoferrin), heme or heme-containing proteins (e.g., hemoglobin or hemopexin) and endogenous or
exogenous siderophores. These iron sources are recognized by specific surface receptors on the OM or CM of Gram-negative or Gram-
positive bacteria, respectively. In Gram-positive bacteria, Fe(III) or Fe(III)-chelator complexes are transported into the cytosol by ABC
ho

transporters coupled to specific CM-anchored lipoprotein receptors. In Gram-negative bacteria, OM receptors transport Fe(III) or Fe(III)-
chelator complexes into the PS thanks to the interaction with the Ton complex (TonB, ExbB, ExbD), which transduces the energy of the
proton gradient on the CM into conformational changes of the receptors. Once in the PS, Fe(III), heme and Fe(III)-loaded siderophores
(ferrisiderophores) are sequestered by specific periplasmic binding proteins that deliver them to cognate ABC transporters (functionally
homologous to those of Gram-positive bacteria). The soluble form Fe(II) can plausibly cross the OM of Gram-negative bacteria through
ut

porins, and is then translocated to the cytosol by the Fe(II) transporter FeoB, which is also present in some Gram-positive bacteria. In
the presence of sufficient amounts of intracellular iron, Fe(II) binds Fur-like repressor proteins (or DtxR-like proteins in some Gram-
positive bacteria). Fe(II)–Fur directly or indirectly represses transcription of genes involved in iron uptake that, therefore, are expressed
only under iron-depleted conditions, in other words, when Fur is mainly present in its apo form (⊥ indicates repression). Fe(II)–Fur also
A

represses the expression of specific small RNAs (e.g., RyhB in Escherichia coli) that inhibit translation of mRNAs encoding a number of
iron-using proteins. As a result, these proteins are mainly expressed under iron-replete conditions.
CM: Cytoplasmic; OM: Outer membrane; PS: Periplasmic space.

the presence of very high Ga(III) concentrations Genome sequencing revealed 64 nonsilent point
(up to 1 mM) display enhanced reactive oxygen mutations in this mutant, mostly occurring in
species (ROS) production and increased expres- genes involved in metabolism, membrane trans-
sion of ROS-detoxifying enzymes, suggesting port, cell division, DNA recombination/repair,
that Ga(III) toxicity is not only due to perturba- signal transduction, energy production and
tion of iron metabolism, but also to the genera- synthesis of siderophores and phenazines [16] .
tion of oxidative stress [11,14] . Recently, a spon- Since several cellular processes were found to
taneous mutant of P. aeruginosa with 12-fold be involved in adaptation/resistance to Ga(III),
increased resistance to Ga(III) was obtained this study suggests that Ga(III) has multiple
through sequential subculturing in the presence targets and resistance mechanisms in bacterial
of an inhibitory concentration of Ga(NO3)3 [16] . cells. The effect of Ga(III) on the expression

future science group www.futuremedicine.com 381


Review  Minandri, Bonchi, Frangipani, Imperi & Visca

of P. aeruginosa iron-repressible virulence fac- Thus, at least in E. coli, Ga(III) activity depends
tors (i.e., factors induced by iron limitation) is on its active transport into the cytoplasm, and
still controversial, and molecular mechanisms Fe(III) uptake mechanisms can contribute to
of interference with gene regulation remain this process. Accordingly, it has been shown
obscure [26,43] . that P. aeruginosa transposon mutants defec-
Irrespective of the mechanism of action, tive in the periplasmic iron-binding protein
Ga(III) needs to enter bacterial cells to reach its HitA or in the cognate inner membrane per-
molecular targets. However, the way by which mease HitB display a three- to fourfold increase
Ga(III) is internalized by bacteria is unclear. in resistance to Ga(III) and >tenfold decrease
Ga(III) has no known role in microbial physi- of the Ga(III) intracellular content [16] , further
ology and biological systems do not discriminate confirming that Ga(III) can exploit endogenous
between Fe(III) and Ga(III), hence the most iron-uptake mechanisms to reach the cytoplasm
plausible hypothesis is that Ga(III) crosses the of bacterial cells.
cytoplasmic membrane of bacteria by exploit-
ing Fe(III)-uptake routes, rather than through ●●Iron chelators as Trojan horses for gallium
delivery to bacterial cells

of
dedicated transport system(s). However, Ga(III)
is efficiently acquired by both wild-type and Recently, there has been a renewed interest in
siderophore-defective P. aeruginosa mutants, the use of natural or synthetic Fe(III) chelators
even under iron-replete conditions that repress as antimicrobial agents [46] . By competing with

ro
the expression of Fe(III)-uptake genes [26] , sug- siderophores for Fe(III) binding, exogenous che-
gesting that siderophore-mediated transport lators can deprive bacterial cells of essential iron.
is not the main route for Ga(III) entrance in This is not a general rule, however, since many
P. aeruginosa. Unfortunately, this work did not bacterial species have evolved the ability to use
assess the actual distribution of cell-associated exogenous chelators as vehicles to deliver Fe(III)
rP
Ga(III) between the periplasmic and cytoplas- to the cell. A paradigmatic example is provided
mic compartments [26] , thus the possibility exists by the iron chelator deferoxamine B (DFO,
that the amount of Ga(III) measured in P. aer- Desferal®, Novartis Pharma, Switzerland),
uginosa cells does not represent the fraction that which can be used as an exogenous iron carrier
actually reaches the cytosol. by several bacteria, including P. aeruginosa [47,48] .
ho

It can be supposed that Ga(III) can cross Although the route(s) by which Ga(III) enters
the outer membrane of Gram-negative bacte- the bacterial cell is uncertain, several studies have
ria by passive diffusion through porins, as also explored the possibility of using chelators from
suggested in the case of soluble Fe(II) ion [38] . both microbial and chemical sources as Trojan
However, the impermeability of the cytoplasmic horses to enhance Ga(III) delivery to bacteria.
membrane of both prokaryotic and eukaryotic This strategy has recently been revisited, and
ut

cells to ions suggests that Ga(III) needs specific several Ga(III)–siderophore and Ga(III)–heme
transport mechanism(s) to reach the cytoplasm. complexes have been evaluated in vitro as new
It has indeed been shown, by small-angle x-ray anti-infectives [49] .
scattering and transmission electron micros- Almost all studies on siderophore-mediated
A

copy, that Ga(III) mainly accumulates in the Ga(III) delivery have been performed on P. aer-
cell-envelope compartment (outer membrane, uginosa as model organism. P. aeruginosa is a par-
periplasm and inner membrane) of P. fluores- adigmatic example of iron-uptake capability, and
cens cells exposed to millimolar concentrations iron metabolism has recently been exploited as
of Ga(III) [44] , confirming that Ga(III) trans- an ideal target for new anti-P. aeruginosa drugs
port across the cytoplasmic membrane is the [22,26,29,50,51] . Banin et al. demonstrated that a
rate-limiting step for internalization by Gram- complex between Ga(III) and the heterologous
negative bacteria. In Escherichia coli expressing siderophore DFO was slightly more effective
the Haemophilus influenzae heterologous pro- than Ga(III) alone in killing P. aeruginosa cells,
teins FbpABC, corresponding to a periplasmic both in established biofilms and in vivo, even
Fe(III)-binding protein (FbpA) and the cognate though the in vitro minimum inhibitory concen-
inner membrane ABC transporter (FbpBC) for trations (MICs) of DFO–Ga(III) and Ga(III)
delivery of Fe(III) from the periplasm to the alone were similar [22] . Although it has been
cytoplasm, both internalization of Ga(III) and demonstrated that DFO receptors present on
susceptibility were significantly increased [45] . the P. aeruginosa outer membrane were required

382 Future Microbiol. (2014) 9(3) future science group


Promises & failures of gallium as an antibacterial agent  Review

for antibacterial activity, the uptake process bacteria, suggesting that the main target(s) of
and the fate of the DFO–Ga(III) complex in Ga(III)–PPIX could be located on the peri-
P. aeruginosa cells were not investigated [22] . The plasmic (extracellular) side of the cytoplasmic
anti-P. aeruginosa activity of Ga(III) complexed membrane [58] . Different from what is observed
with either DFO or the iron chelator citrate has with Ga(III) alone, the inhibitory activity of the
also been compared [29] . The Ga(III)–citrate Ga(III)–PPIX complex is not significantly coun-
complex showed higher bacteriostatic and bac- teracted by an excess of Fe(III) [30] . Moreover,
tericidal activities than DFO–Ga(III), and was obligate and facultative anaerobic bacteria using
also more effective in protecting macrophages fermentation for energy production were found
from infection, plausibly due to a more efficient to be resistant to Ga(III)–PPIX, while E. coli
uptake of Ga(III)–citrate by P. aeruginosa cells mutants impaired in ROS detoxification mecha-
[29] . However, the activities of Ga(III)–citrate nisms were hypersensitive to this compound [30] .
and Ga(III) alone were not compared in this Altogether, these observations led to the hypoth-
study, hampering the direct evaluation of the esis that the inhibitory activity of Ga(III)–PPIX
effect of chelator-mediated transport of Ga(III) does not rely on intracytoplasmic release of free

of
on bacterial growth. Notably, Ga(III)–citrate Ga(III), but on insertion of Ga(III)–PPIX into
was found to exert a negligible inhibitory activ- membrane-bound cytochromes, with conse-
ity on a number of other species, including E. quent disruption of the electron flow, ROS
coli [29] . This is of particular interest considering overproduction and oxidative stress [58] .

ro
that a functional Fe(III)-citrate transport system
is present in E. coli K-12 [52] , and indicates that In vitro & in vivo antibacterial activity of
citrate-mediated active uptake of Ga(III) may gallium compounds
not be sufficient for Ga(III)–citrate toxicity. The interest in Ga(III) as an antibacterial agent
Different from heterologous siderophores or has been resuscitated by the paucity of effec-
rP
chemical chelators, endogenous siderophores tive antibiotics to combat emerging multid-
have no apparent effect on Ga(III) activity in rug-resistant (MDR) bacterial pathogens (i.e.,
P. aeruginosa. Although both Ga(III)–pyo- isolates nonsusceptible to at least one agent in
verdine and Ga(III)–pyochelin complexes are three or more antimicrobial categories accord-
recognized and actively internalized by cognate ing to [59]). Several investigations indicate that
ho

outer membrane receptors [53,54] , the inhibitory Ga(III) is able to inhibit bacterial growth both
activity of Ga(III) on P. aeruginosa was not in vitro and in vivo [20–30,56], prompting clinical
influenced by the presence of these siderophores investigations of the efficacy of Ga(III)-based
and their uptake systems [26,53,54] . However, the antimicrobial chemotherapy.
anti-P. aeruginosa effect of preformed Ga(III)–
pyoverdine and Ga(III)–pyochelin complexes ●●Activity on planktonic bacterial pathogens
ut

has neither been tested nor compared with that Most pathogenic bacteria are able to multiply
of Ga(III) alone. Thus, the few data available in biological fluids and/or in the environment
indicate that Ga(III) complexation with sidero- as free planktonic cells, a mode of growth that
phores and/or iron chelators does not signifi- is mimicked by cultures in liquid media, as
A

cantly increase its antibacterial activity, suggest- for antimicrobial susceptibility tests. Several
ing that the internalization process is not the Ga(III) compounds have been tested against
limiting step for the activity of Ga(III). a variety of bacterial species in liquid cultures,
Promising results were obtained with the under an array of conditions that often hamper
Ga(III)–protoporphyrin IX (PPIX or apo- the direct comparison of results from different
heme) complex. It has been demonstrated that sources. Hereafter, the most relevant data will
Ga(III)–PPIX has potent antibacterial activ- be discussed.
ity against a number of Gram-negative and Two independent studies have recently dem-
Gram-positive bacteria [30,49,55–57] . Although onstrated strong activity of Ga(NO3)3 on the
Ga(III) –PPIX can exploit heme receptors nosocomial pathogen Acinetobacter bauman-
for transport across the outer membrane of nii [20,60] . Low concentrations were effective
Gram-negative bacteria [57,58] , heme uptake in inhibiting the growth of both laboratory
systems on the cytoplasmic membrane are not and MDR A. baumannii isolates in iron-poor
required for Ga(III)–PPIX-mediated inhibi- media (inhibitory concentrations by 90%
tion of both Gram-negative and Gram-positive [IC 90 s] ranged from 2 to 80 μM) (Table 1) ,

future science group www.futuremedicine.com 383


Review  Minandri, Bonchi, Frangipani, Imperi & Visca

Table 1. In vitro antibacterial activity of gallium.


Effect on Microorganism Ga(III) Growth condition (medium)§ Compound inhibitory Ga(III) concentration Ref.
compound concentration (μM)†
Planktonic Acinetobacter Ga(NO3)3 Chemically-undefined medium IC90 >100 μM >100 [60]
growth baumannii (10% TSB)
Chemically-defined medium IC90 = 100 μM 100 [60]
(RPMI 1640)
Chemically-defined medium IC90 = 3.2 μM 3.2 [60]
(RPMI + 10% serum)
Chemically-defined medium (M9- IC90 = 2–80 μM 2–80 [20]
DIP)
Chemically-undefined medium IC90 = 4–64 μM 4–64 [20]
(human serum)
Bacillus subtilis Ga(III)–PPIX Chemically-undefined medium MIC = 0.2 μg/ml 0.3 [30]
(NB)

of
Burkolderia cepacia Ga(NO3)3 Chemically-defined medium MIC = 250 μM 250 [28]
complex (CDM no iron)
Enterococcus Ga(III)–Cit Chemically-undefined medium IC90 = 1280–2560 μM 1280–2560 [29]
faecalis (Iso-Sensitest; Fisher Scientific,
UK)

Escherichia coli
DFO–Ga(III)

Ga(III)–Cit,
DFO–Ga(III) ro
Chemically-undefined medium
(Iso-Sensitest)
Chemically-undefined medium
(Iso-Sensitest)
IC90 = 1280–5120 μM

IC90 = 2560–5120 μM
1280–5120

2560–5120
[29]

[29]
rP
Ga(III)–Cit, Chemically-defined medium IC90 = 1280–5120 μM 1280–5120 [29]
DFO–Ga(III) (HHW)
Ga(III)–PPIX Chemically-undefined medium MIC <0.5 μg/ml <0.8 [30]
(NB + DIP)
Helicobacter pylori Ga(III)–PPIX Chemically-undefined medium MIC = 0.19 μg/ml 0.3 [30]
(Brucella broth + 2% serum)
ho

Klebsiella Ga(III)–Cit Chemically-defined medium IC90 = 1280–5120 μM 1280–5120 [29]


pneumonia (HHW)
Chemically-undefined medium IC90 = 1280–5120 μM 1280–5120 [29]
(Iso-Sensitest)
DFO–Ga(III) Chemically-defined medium IC90 = 1280–2560 μM 1280–2560 [29]
(HHW)
ut

Chemically-undefined medium IC90 = 2560–5120 μM 2560–5120 [29]


(Iso-Sensitest)
Ga(III)–PPIX Chemically-undefined medium MIC = 2 μg/ml 3.2 [30]
(NB)
A

Lysteria Ga(III)–PPIX Chemically-undefined medium MIC = 0.2 μg/ml 0.3 [30]


monocytogenes (NB)
Mycobacterium Ga(NO3)3 Chemically-undefined medium IC50 ≈ 250 μM‡ 250 [27]
avium complex (BACTEC™; BD, NJ, USA)
M. avium subsp. Ga(NO3)3 Chemically-undefined medium IC90 = 200–743 μM 200–743 [62]
paratuberculosis (BACTEC)
Mycobacterium Ga(III)–PPIX Chemically-undefined medium MIC = 0.4 μg/ml 0.6 [30]
bovis (NB)

Actual concentration of the Ga3+ ion after complex dissociation.

IC50s were not clearly stated, but extrapolated from the references.
§
Refer to citations for composition of culture media.
7H9: Middlebrook 7H9 broth; BACTEC: BACTEC 12b broth; BBC: Biofilm bactericidal concentration; BIC: Biofilm inhibitory concentration; BM2 succinate: BM2 minimal medium
supplemented with 34 mM succinate; CDM: Chemically-defined medium; DFO–Ga(III): Desferoxamine–gallium; DIP: 2’2-dipyridyl; Ga(III)–PPIX: Gallium–protoporphyrin
IX; Ga(III)–Cit: Gallium–citrate; GaM: Gallium maltolate; HHW: Hussain, Hastings & White medium; IC50: Compound concentrations that inhibit growth by 50%; IC90: Compound
concentrations that inhibit growth by 90%; LB: Luria-Bertani broth; M9-DIP: Mimimal medium 9 supplemented with 0.1 mM 2,2’-dipyridyl; MBC: Minimum bactericidal
concentration; MBBC: Minimum biofilm bactericidal concentration; MBIC: Minimum biofilm inhibitory concentration; MIC: Minimum inhibitory concentration (lowest
concentration causing no visible growth); MM: Minimal medium; NB: Nutrient broth; RPMI: Roswell Park Memorial Institute medium; Tf–Ga(III): Transferrin–gallium; TSB: Tryptic soy
broth.

384 Future Microbiol. (2014) 9(3) future science group


Promises & failures of gallium as an antibacterial agent  Review

Table 1. In vitro antibacterial activity of gallium (cont.).


Effect on Microorganism Ga(III) Growth condition (medium)§ Compound inhibitory Ga(III) concentration Ref.
compound concentration (μM)†
Planktonic Mycobacterium Ga(III)–PPIX Chemically-undefined medium MIC = 0.4 μg/ml 0.6 [30]
growth smegmatis (LB)
(cont.)
Mycobacterium Ga(NO3)3 Chemically-undefined medium IC50 ≈ 60–250 μM‡ 60–250 [27]
tuberculosis (BACTEC)
Chemically-defined medium IC50 = 1.25–2.50 μM 1.25–2.50 [27]
(7H9)
Neisseria spp. Ga(III)–PPIX Chemically-undefined medium MIC = 0.2 μg/ml 0.3 [30]
(NB)
Proteus mirabilis Ga(III)–Cit, Chemically-undefined medium IC90 = 1280 μM 1280 [29]
DFO–Ga(III) (Iso-Sensitest)
Ga(III)–Cit, Chemically-defined medium IC90 = 1280 μM 1280 [29]

of
DFO–Ga(III) (HHW)
Pseudomonas Ga(NO3)3 Chemically-defined medium IC90 = 1–40 μM 1–40 [26]
aeruginosa (BM2 succinate)
Chemically-defined medium MIC = 62.5 μM 62.5 [28]

ro
(CDM no iron)
GaCl3, DFO– Chemically-undefined medium MIC = 32 μM 32 [22]
Ga(III) (1% TSB)
Ga(III)–Cit Chemically-defined medium IC90 = 20–40 μM 20–40 [29]
(HHW)
rP
Chemically-undefined medium IC90 = 80–640 μM 80–640 [29]
(Iso-Sensitest)
DFO–Ga(III) Chemically-defined medium IC90 = 40–640 μM 40–640 [29]
(HHW)
Chemically-undefined medium IC90 = 160–1280 μM 160–1280 [29]
(Iso-Sensitest)
ho

Rhodococcus equi Ga(NO3)3 Chemically-defined medium MBC = 50 μM 50 [24]


(MM)
GaM Chemically-defined medium MIC90 = 8 μM 8 [69]
(minimal medium without iron)
Staphylococcus Ga(III)–Cit Chemically-defined medium IC90 = 320–5120 μM 320–5120 [29]
ut

aureus (HHW)
Chemically-undefined medium IC90 = 2560–5120 μM 2560–5120 [29]
(Iso-Sensitest)
DFO–Ga(III) Chemically-defined medium IC90 = 320–5120 μM 320–5120 [29]
(HHW)
A

Chemically-undefined medium IC90 = 2560–5120 μM 2560–5120 [29]


(Iso-Sensitest)
GaM Chemically-defined medium MIC = 375–2000 μg/ml 847–4494 [21]
(RPMI)
Chemically-defined medium MIC = 0.06–4.00 mg/ml 140–8988 [64]
(RPMI 1640)
Ga(III)–PPIX Chemically-undefined medium MIC = 1.0–2.5 μg/ml 1.6–4.0 [30]
(NB)

Actual concentration of the Ga3+ ion after complex dissociation.

IC50s were not clearly stated, but extrapolated from the references.
§
Refer to citations for composition of culture media.
7H9: Middlebrook 7H9 broth; BACTEC: BACTEC 12b broth; BBC: Biofilm bactericidal concentration; BIC: Biofilm inhibitory concentration; BM2 succinate: BM2 minimal medium
supplemented with 34 mM succinate; CDM: Chemically-defined medium; DFO–Ga(III): Desferoxamine–gallium; DIP: 2’2-dipyridyl; Ga(III)–PPIX: Gallium–protoporphyrin
IX; Ga(III)–Cit: Gallium–citrate; GaM: Gallium maltolate; HHW: Hussain, Hastings & White medium; IC50: Compound concentrations that inhibit growth by 50%; IC90: Compound
concentrations that inhibit growth by 90%; LB: Luria-Bertani broth; M9-DIP: Mimimal medium 9 supplemented with 0.1 mM 2,2’-dipyridyl; MBC: Minimum bactericidal
concentration; MBBC: Minimum biofilm bactericidal concentration; MBIC: Minimum biofilm inhibitory concentration; MIC: Minimum inhibitory concentration (lowest
concentration causing no visible growth); MM: Minimal medium; NB: Nutrient broth; RPMI: Roswell Park Memorial Institute medium; Tf–Ga(III): Transferrin–gallium; TSB: Tryptic soy
broth.

future science group www.futuremedicine.com 385


Review  Minandri, Bonchi, Frangipani, Imperi & Visca

Table 1. In vitro antibacterial activity of gallium (cont.).


Effect on Microorganism Ga(III) Growth condition (medium)§ Compound inhibitory Ga(III) concentration Ref.
compound concentration (μM)†
Planktonic Staphylococcus Ga(III)–Cit Chemically-defined medium IC90 = 1280–2560 μM 1280–2560 [29]
growth epidermidis (HHW)
(cont.)
Chemically-undefined medium IC90 = 2560–5120 μM 2560–5120 [29]
(Iso-Sensitest)
DFO–Ga(III) Chemically-defined medium IC90 = 640–2560 μM 640–2560 [29]
(HHW)
Chemically-undefined medium IC90 = 1280–5120 μM 1280–5120 [29]
(Iso-Sensitest)
GaM Chemically-defined medium MIC = 100–200 μg/ml 225–450 [21]
(RPMI)
Staphylococcus GaM Chemically-defined medium MIC = 0.25–4.00 mg/ml 562–8988 [64]

of
pseudintermedius (RPMI 1640)
Vibrio cholerae Ga(III)–Cit Chemically-defined medium MIC = 1280–2560 μM 1280–2560 [29]
(HHW)
DFO–Ga(III) Chemically-defined medium MIC = 1280 μM 1280 [29]

ro
(HHW)
Ga(III)–Cit, Chemically-undefined medium MIC = 1280 μM 1280 [29]
DFO–Ga(III) (Iso-Sensitest)
Yersinia Ga(III)–PPIX Chemically-undefined medium MIC = 0.4 μg/ml 0.6 [30]
enterocolitica (NB + DIP)
rP
Yersinia Ga(III)–PPIX Chemically-undefined medium MIC = 0.4 μg/ml 0.6 [30]
pseudotuberculosis (NB + DIP)
Biofilm P. aeruginosa Ga(NO3)3 Flow cells (1% TSB) BIC = 0.5 μM 0.5 [26]
BBC = 100 μM 100
GaCl3, DFO– Flow cells (1% TSB) BIC = 1 μM 1 [22]
Ga(III)
ho

DFO–Ga(III) Flow cells (1% TSB) BBC = 1 mM 1000 [22]


Ga(III)–Cit Microtitre plates MBIC = 10 μM 10 [29]
(20% Iso-Sensitest)
B. cepacia complex Ga(NO3)3 Microtitre plates (CDM no iron) MBBC = 32 mg/l 125 [28]
S. aureus GaM Microtitre plates (RPMI) MBIC = 3000–6000 6741–13482 [21]
ut

μg/ml
S. epidermidis GaM Microtitre plates (RPMI) MBIC = 94–3000 μg/ml 211–6741 [21]
Intracellular M. tuberculosis Ga(NO3)3 Human macrophages IC50 = 30–100 μM 30–100 [27]
growth
A

MBC = 500 μM 500
Tf–Ga(III) Human macrophages IC50 = 62.5 μM 125.0 [27]

Actual concentration of the Ga3+ ion after complex dissociation.

IC50s were not clearly stated, but extrapolated from the references.
§
Refer to citations for composition of culture media.
7H9: Middlebrook 7H9 broth; BACTEC: BACTEC 12b broth; BBC: Biofilm bactericidal concentration; BIC: Biofilm inhibitory concentration; BM2 succinate: BM2 minimal medium
supplemented with 34 mM succinate; CDM: Chemically-defined medium; DFO–Ga(III): Desferoxamine–gallium; DIP: 2’2-dipyridyl; Ga(III)–PPIX: Gallium–protoporphyrin
IX; Ga(III)–Cit: Gallium–citrate; GaM: Gallium maltolate; HHW: Hussain, Hastings & White medium; IC50: Compound concentrations that inhibit growth by 50%; IC90: Compound
concentrations that inhibit growth by 90%; LB: Luria-Bertani broth; M9-DIP: Mimimal medium 9 supplemented with 0.1 mM 2,2’-dipyridyl; MBC: Minimum bactericidal
concentration; MBBC: Minimum biofilm bactericidal concentration; MBIC: Minimum biofilm inhibitory concentration; MIC: Minimum inhibitory concentration (lowest
concentration causing no visible growth); MM: Minimal medium; NB: Nutrient broth; RPMI: Roswell Park Memorial Institute medium; Tf–Ga(III): Transferrin–gallium; TSB: Tryptic soy
broth.

and the Ga(III) effect was gradually reversed Ganite regimens [20] . These findings represent
by increasing FeCl 3 concentration [20,60] . an important advance in antibacterial chemo-
Interestingly, Ga(NO3)3 was even more effec- therapy, given the paucity of antibiotics retain-
tive when A. baumannii was grown in human ing activity against A. baumannii, and the risk
serum, showing IC90 values within the range of posed by the emergence and dissemination
Ga(III) plasma levels achievable with standard of pandrug-resistant (PDR) clones [61] (i.e.,

386 Future Microbiol. (2014) 9(3) future science group


Promises & failures of gallium as an antibacterial agent  Review

nonsusceptibile to all agents in all antimicrobial very high MICs (in the mM range) were observed
categories according to [59]). with both complexes for all but one species tested
Ga(III) activity was investigated in-depth (Table 1) [29] . P. aeruginosa was the most sensitive
on planktonic cells of the opportunistic patho- to Ga(III)–citrate (IC90 range 20–40 μM) and
gen P. aeruginosa [22,26] . Ga(NO3)3 was highly DFO–Ga(III) (IC90 range 40–640 μM) in mini-
effective in inhibiting the growth of clini- mal medium [29] . However, these IC90 values are
cal P. aeruginosa isolates, showing IC90 values higher than those observed with Ga(NO3)3 or
between 1 and 40 μM, and in this case Fe(III) GaCl3 alone [22,26] .
also counteracted Ga(III) activity [26] . Similar Gallium maltolate (GaM, 1:3 gallium:maltol
inhibitory activity was reported using gallium molar ratio) is an orally active Ga(III) compound
chloride (GaCl3) on the reference P. aeruginosa endowed with high bioavailablity and low toxic-
strain PAO1, with MIC = 32 μM (Table 1) [22] . ity [63] , whose activity has been assessed against
Moreover, Ga(NO3)3 exerted a dose-dependent different Staphylococcus species [21,64] . In general,
bactericidal effect on P. aeruginosa starting from Staphylococcus aureus showed poor susceptibil-
concentrations as low as 10 μM [26] . ity to GaM, as also reported for Ga(III)–cit-

of
These data suggest that existing resistance to rate (Table 1) [21,29] . GaM activity was also dis-
conventional antibiotics does not compromise couraging for Staphylococcus pseudintermedius
the activity of Ga(III) against MDR pathogens veterinary isolates (Table 1) [64] .
like A. baumannii and P. aeruginosa [20,26] . GaM was tested on Salmonella spp. with the

ro
Ga(NO3)3 activity was also investigated in the purpose of reducing Salmonella fecal shedding
Burkholderia cepacia complex species, but the in cattle [25] . The growth of Salmonella enterica
MICs were 250 μM for all species tested (Table 1) serovar Newport was inhibited by GaM in a
[28] . Intriguingly, P. aeruginosa PAO1 grown in time- and dose-dependent manner, albeit a sig-
the same medium as an internal control, showed nificant effect was evident only at high GaM
rP
a MIC much higher (62.5 μM) than previously concentrations (250–500 μM) in minimal
reported inhibitory concentrations [22,26] , pre- growth medium [25] .
sumably owing to different composition and As anticipated , Ga(III)–PPIX is endowed
iron content of the media. with strong antibacterial activity against a wide
The ability of Ga(NO3 ) 3 to inhibit the range of Gram-positive and Gram-negative
ho

planktonic growth of mycobacteria has been bacterial pathogens, even in iron-sufficient


investigated on the Mycobacterium avium com- growth media [30,49,55,56] . Among the bacte-
plex, M. avium subsp. paratuberculosis (MAP) ria tested, Bacillus subtilis, E. coli, Helicobacter
and Mycobacterium tuberculosis strains. High pylori, Listeria monocytogenes, Mycobacterium
Ga(NO3 ) 3 concentrations (>200 μM) were bovis, Neisseria gonorrhoeae, Neisseria menin-
needed to reduce planktonic growth of M. gitidis, Yersinia enterocolitica and Yersinia pseu-
ut

avium complex and MAP strains, likely as a dotuberculosis, were the most sensitive [MICs ≤
consequence of the high iron content of the 0.5 μg/ml, corresponding to 0.8 μM Ga(III)]
media used (Table 1) [27,62] . In fact, the Ga(NO3)3 (Table 1) [30] . Good Ga(III)–PPIX inhibitory
concentration required to inhibit M. tuberculo- activity was also recently reported for A. bau-
A

sis planktonic growth in a chemically-defined mannii [49] . Conversely, some bacterial species
iron-poor medium was much lower (concentra- were found to be resistant to Ga(III)–PPIX,
tion that inhibited the growth by 50% [IC50] and these were Gram-negative species lack-
was ∼2.0 μM) (Table 1) [27,62] . ing heme uptake systems (S. enterica serovar
In addition to inorganic Ga(III) salts, a num- Typhimurium) and/or microorganisms that
ber of Ga(III) complexes have been tested on do not contain cytochromes such as facultative
bacterial pathogens with the aim of enhancing anaerobes (Streptococcus pyogenes and Enterococcus
the antimicrobial efficacy of Ga(III). faecalis) [30] .
As mentioned above, although P. aeruginosa
uses DFO as an iron carrier, DFO–Ga(III) (1:1 ●●Activity on bacterial biofilms
molar ratio) was bacteriostatic and bactericidal at Bacterial cells living in biofilms are more resist-
almost the same concentrations as GaCl3 alone ant to antibiotics than their planktonic counter-
[22] . The antibacterial activity of Ga(III)–citrate part, mostly owing to the composition and low
(1:2 molar ratio) and DFO–Ga(III) complexes permeability of the extracellular matrix, which
was also compared in another study; however, makes biofilms often responsible for persistent

future science group www.futuremedicine.com 387


Review  Minandri, Bonchi, Frangipani, Imperi & Visca

or chronic infections. Given the favorable diffu- reduction of M. tuberculosis growth at 62.5
sion properties of Ga(III) complexes, a number μM (corresponding to 125 μM Ga(III)) [27] ,
of studies have exploited the possibility of using comparable with the effect of Ga(NO3)3 alone.
Ga(III) as a biofilm-perturbing agent (Table 1) . Growth inhibition >70% was achieved at
While P. aeruginosa biofilms are particularly Tf–Ga(III) or Lf–Ga(III) concentrations as low
difficult to eradicate by means of antibiotic as 5 μM (corresponding to 10 μM of Ga(III))
therapy, they were found to be highly sensi- for Francisella tularensis and Francisella novicida
tive to Ga(III). It was shown that a sub-MIC strains [extrapolated from [66] , concomitant with
concentration of Ga(NO3)3 strongly inhibited decreased activity of key iron-containing antiox-
biofilm formation by P. aeruginosa (Table 1) , and idant enzymes, such as catalase and superoxide
that inhibition was reversed by biofilm perfusion dismutase [66] . Within macrophages, F. tula-
with FeCl3 [26] . Impressively, 100 μM Ga(NO3)3 rensis was inhibited by 100 μM Tf–Ga(III)
killed P. aeruginosa cells deeply embedded in the or Lf–Ga(III) (corresponding to 200 μM
biofilm matrix, where most antibiotics lose their Ga(III)), although the activity of Tf–Ga(III)
efficacy due to poor permeation (Figure 2) [26] . and Lf–Ga(III) were not compared with that of

of
Other Ga(III) compounds, like DFO–Ga(III), Ga(III) alone [66] .
GaCl 3, and Ga(III)–citrate, showed potent Potent inhibitory activity on the extracellular
suppressive activity on P. aeruginosa biofilm planktonic growth of R. equi was reported using
formation in the micromolar range [22,29] . both GaM (IC90 = 8 μM) [69] and Ga(NO3)3
(bactericidal effect at 50 μM) [24] . Accordingly,

ro
Similar effects were also observed using 64 mg/l
Ga(NO3)3 (corresponding to 250 μM Ga(III)) treatment of murine macrophage-like cells with
on preformed biofilms of bacteria belonging to 10 μM GaM caused >threefold decrease in
the B. cepacia complex [28] . intracellular R. equi concentrations [70] .
Consistent with the poor activity of Ga(III) Finally, Ga(III)–PPIX showed potent inhi-
rP
on planktonic staphylococcal cells, the minimal bition of cell invasion and intracellular multi-
biofilm inhibitory concentration of GaM was plication of the oral pathogen Porphyromonas
very high (mM range) for both laboratory and gingivalis [56] .
clinical Staphylococcus strains (Table 1) [21] . As a whole, these results (Table 1) demonstrate
that Ga(III) affects the growth of relevant
ho

●●Activity on intracellular bacterial bacterial pathogens in different cellular models.


pathogens
Several pathogenic bacteria such as M. tubercu- ●●Protective activity of gallium in animal
losis, Francisella spp. and Rhodococcus equi have models of bacterial infection & veterinary
evolved the ability to infect and replicate in a applications
variety of human cells, especially macrophages, The majority of animal studies on the in vivo
ut

where intracellular iron is essential for their life efficacy of Ga(III) compounds were conducted
[27,65,66] . Bacterial iron-uptake systems also play in mice infected by different routes with a given
a role intracellularly, allowing iron retrieval from bacterial species, and then treated with Ga(III)
Tf and Lf within macrophages [27] . At the site of compounds. In addition, for in vivo studies,
A

infection, iron uptake by activated macrophages direct comparison of outcomes is complicated


prevalently occurs via specific Lf and Tf recep- by differences in the type of infection, thera-
tors, and therefore Tf–Fe(III) and Lf–Fe(III) peutic dosage, and routes of Ga(III) administra-
complexes represent an important component tion. Most remarkably, the relevance for human
of the macrophage intracellular iron pool [67,68] . applications, that is the translation from human
Ga(NO3)3 treatment of human macrophages therapeutic dose to animal dose, has rarely been
infected with both antibiotic-susceptible and addressed.
MDR M. tuberculosis resulted in dose-depend- The earliest evidence of the in vivo antimicro-
ent growth inhibition, with IC50s ranging from bial efficacy of Ga(III) dates back to 1931, when
ca. 30 to ca. 100 μM, depending on the strain Levaditi et al. demonstrated a protective effect
(extrapolated from [27]). Moreover, Ga(NO3)3 of Ga(III)–tartrate from Treponema pallidum
exerted a bactericidal effect on M. tuberculosis or Trypanosoma evansi infections in different
within macrophages at a concentration of 500 animal models (Table 2) [71] .
μM, [27] . Challenge of infected macrophages More recently, the in vivo activity of Ga(NO3)3
with Tf–Ga(III) (molar ratio 1:2) showed >50% against A. baumannii was investigated in

388 Future Microbiol. (2014) 9(3) future science group


Promises & failures of gallium as an antibacterial agent  Review

0 µM Ga(III) 0.5 µM Ga(III)

of
24 h 48 h 72 h

Cross section
ro
rP
Side view
ho

Figure 2. Effect of Ga(III) on Pseudomonas aeruginosa biofilms. (A) Ga(III) prevents Pseudomonas
aeruginosa biofilm formation. Confocal microscopic images of green-fluorescent protein-labeled
P. aeruginosa in flow cells perfused with growth medium without (left) or with (right) Ga(NO3)3 at the
ut

indicated concentration, 5 days after inoculation. (B) Ga(III) kills established biofilm. Biofilm of GFP-
labeled P. aeruginosa were grown for 3 days in the absence of Ga(NO3)3 and then treated with 100 μM
Ga(NO3)3 for the indicated time period. Cells were stained with propidium iodide (30 μM) to label
dead cells. Images in (A) and (B, top) are top-down views (x–y plane); (B, bottom) images are side
A

views (x–z plane); scale bars: 50 μm.


Adapted with permission from [26].
For color images please see www.futuremedicine.com/doi/full/10.2217/fmb.14.3

two different models, the insect Galleria mel- for treatment of patients infected with PDR
lonella and the neutropenic mouse [20,60] . A. baumannii.
Administration of Ga(NO3 ) 3, at concentra- Kaneko et al. tested the antibacterial activ-
tions mimicking the human therapeutic dose, ity of Ga(NO3)3 against P. aeruginosa in both
protected G. mellonella from A. baumannii- acute and chronic mouse lung infection. In the
mediated killing [20] . Coherently, intraperito- acute model, fatal sepsis due to intratracheal
neal administration of Ga(NO3)3 was effective infection with a lethal dose of P. aeruginosa
in reducing A. baumannii load in the lungs of was prevented by a single intranasal adminis-
neutropenic mice [60] . Putting together in vitro tration of Ga(NO3)3 3 h postinfection (Table 2)
and in vivo results (Tables 1 & 2) , it can be con- [26] . As expected, intranasal administration of
cluded that Ga(NO3)3 holds promise as a new Fe(III) prior to infection abrogated Ga(III)
anti-A. baumannii agent, raising new hope activity [26] . Similarly, Ga(NO3)3 was effective

future science group www.futuremedicine.com 389


Review  Minandri, Bonchi, Frangipani, Imperi & Visca

Table 2. In vivo antibacterial activity of gallium.


Ga(III) Microorganism Therapeutic regimen (route of Concentration in μmol/kg Type of infection (animal Ref.
compound administration) of single dose model)
Ga(III) tartrate Treponema 30–45 mg/kg (im.) 136–205 Acute infection (rabbit) [71]
pallidum 15 mg/kg (iv.) 68 Acute infection (rabbit) [71]
Trypanosoma 225 mg/kg (iv.) 1024 Acute infection (mice) [71]
evansi
Ga(NO3)3 Acinetobacter Single dose of 1200 μmol/kg 1200 Hemolymph infection [20]
baumannii (hemolymph injection) (Galleria mellonella)
Two doses, prior and after 98 Pulmonary infection [60]
infection, of 25 mg/kg (ip.) (neutropenic mouse)
Pseudomonas Single dose of 50 μl of a 250 mM 600† Acute pulmonary infection [26]
aeruginosa solution (in.) (mouse)
Three daily doses of 50 μl of 600† Chronic airway infection [26]
a 250 mM solution for 3 days model (mouse)

of
(intranasal)
Mycobacterium Two doses, prior and after 78 Intestinal infection (calves) [72]
avium subsp. infection, of 20 mg/kg (oral)
paratuberculosis

ro
Francisella novicida Single dose of 50 μl of 250 mg/ 2330† (single dose) + 40† Pulmonary infection [66]
ml solution (in.) + daily doses of (daily dose) (mouse)
10 mg/kg for 15 days (ip.)
GaM P. aeruginosa Single dose of 25 mg/kg (sc.) 56 Wound infection (burned [23]
mouse)
rP
A. baumannii Single dose of 100 mg/kg (sc.) 225 Wound infection (burned [23]
mouse)
Staphylococcus Single dose of 100 mg/kg (sc.) 225 Wound infection (burned [23]
aureus mouse)
Rhodococcus equi Daily dose, prior and after 22.5 Peritoneum infection [24]
infection, of 10 mg/kg for (mouse)
ho

16 days (oral)
Ga(III)–PPIX Neisseria Single dose prior infection with ND Vaginal infection (mouse) [55]
gonorrhoeae 2 μg (topic)
Values extrapolated considering a mouse average weight of 21 g.

Ga(III)–PPIX: Gallium–protoporphyrin IX; GaM: Gallium maltolate; im.: Intramuscular; in.: Intranasal; ip.: Intraperitoneal; iv., Intravenous; ND: Not determinable; sc.: Subcutaneous.
ut

in a murine model of chronic airway infection, lymph node tissues was significantly decreased
where administration of Ga(NO3)3 starting 5 h (Table 2) [72] .
postinfection, three times/day by inhalation GaM has been evaluated against P. aerugi-
for 9 days, reduced the bacterial load in the nosa, using a thermally injured mouse model of
A

lung (Table 2) [26] . These in vivo results are fully acute infection. Subcutaneously administered
consistent with in vitro studies , reporting high GaM prevented fatal infection, indicating that
susceptibility of P. aeruginosa to Ga(III)-based P. aeruginosa proliferation and systemic spread
compounds. were blocked [23] . The protective activity of GaM
In vivo activity of Ga(NO3)3 has also been in vivo was also observed for A. baumannii and
assayed in a mouse model of F. novicida lung S. aureus burn eschar infection, albeit at higher
infection. Multiple Ga(NO3 ) 3 administra- locally injected doses of GaM than those used
tions protected mice from lethal infection [66] , to suppress P. aeruginosa (Table 2) [23] . Despite
although at relative high dosages (Table 2) . the poor susceptibility reported in vitro (Table 1),
Ga(NO3 ) 3 activity has also been tested in this study holds promises for the use of GaM, a
calves, in order to determine protection from Ga(III)-compound endowed with low toxicity
infection with MAP and assess putative collat- [63] , against S. aureus infection.
eral effects of treatment [72] . No adverse effects Oral GaM reduced R. equi burden in spleen,
were observed in calves orally treated with lung and liver in a mouse model of infection
Ga(NO3)3, whilst MAP burden in intestinal and (Table 2) [24] . The use of GaM in veterinary

390 Future Microbiol. (2014) 9(3) future science group


Promises & failures of gallium as an antibacterial agent  Review

medicine was also extended to cattle experi- To further enhance the delivery of the
mentally infected with S. enterica; however, no Ga(III)–GEN combination, the activity of these
reduction in Salmonella fecal level was observed compounds co-encapsulated in liposomes has
[25] . These results, together with in vitro data been evaluated on clinical P. aeruginosa isolates
(Table 1) , discourage the use of Ga(III) against [76] . The formulation, called Lipo–Ga–GEN,
Salmonella spp. displayed lower MIC and MBC than Lipo–GEN
Ga–PPIX is endowed with strong in vitro alone on planktonic growing P. aeruginosa [76] .
activity against N. gonorrhoeae [30] . Ga–PPIX Moreover, Lipo–Ga–GEN successfully killed
activity was also confirmed in vivo, resulting in P. aeruginosa biofilms, while the correspond-
substantial reduction of gonococcal coloniza- ing concentrations of GEN (8 mg/l), Ga(III)
tion in an experimental murine model of vaginal (0.6 μM), Lipo–Ga (0.6 μM), or Lipo–GEN
infection (Table 2) [55] . (8 mg/l) supplied alone, had much weaker
In conclusion, different animal models effects [76] .
of infection provided evidence of an over- Another example of Ga(III) –antibiotic
all good efficacy and tolerability of Ga(III)- synergism was offered by the high activity of

of
based compounds. The most promising results Ga(NO3)3 in combination with colistin (poly-
were obtained in laboratory animals infected myxin E) [20] , which represents the last therapeu-
with MDR bacteria, such as P. aeruginosa and tic option for PDR A. baumannii. Since colistin
A. baumannii. MDR P. aeruginosa infections acts by perturbating the outer membrane integ-

ro
represent the main cause of death and compli- rity, it has been hypothesized that it could facili-
cation in an array of chronic pulmonary dis- tate Ga(III) diffusion into A. baumannii cells,
eases, thus the possibility of developing new thereby promoting growth inhibition [20] .
Ga(III)-based anti-pseudomonas therapies is
fascinating. The growing interest in this area is Conclusion & future perspective
rP
testified by the recent set up of a Phase 1 clini- The necessity of iron for bacterial growth and
cal trial aimed at monitoring the effect of the the requirement of efficient iron assimilation sys-
FDA-approved Ga(NO3)3 formulation (Ganite) tems for infection have made iron transport and
in cystic fibrosis patients chronically infected by metabolism very attractive targets for the devel-
P. aeruginosa [73] . opment of new antibacterial therapies. Many
ho

efforts have been put forth to assess the drug-


Synergistic activity of Ga(III) with gability of bacterial Fe(III) transport systems;
antimicrobials however, their diversity, functional redundancy
Several studies have explored the activity of and species-specificity have so far precluded
Ga(III) in combination with antibiotics, in the development of broad-spectrum antibacte-
search for useful synergistic effects (i.e., the rial agents targeting iron uptake [46] . On the
ut

reduction in the active concentration of both other hand, the Fe(III)-mimetic properties of
antibiotic and gallium compared with individual Ga(III) provided a shortcut to efficient target-
components used alone, according to [74]). ing of bacterial iron metabolism, circumventing
The antimicrobial efficacy of DFO–Ga(III) the problem of Fe(III) uptake specificity. The
A

in combination with gentamicin (GEN) has repositioning of Ga(III) as an antibacterial agent


been tested on both stationary phase and bio- offered the opportunity to take advantage of the
film-growing P. aeruginosa cells [22] . The DFO– wealth of knowledge on the already established
Ga(III)–GEN combination caused a higher pharmacological properties of Ga(III)-based
reduction of bacterial viability than GEN or formulations, thus accelerating clinical testing
DFO–Ga(III) supplied alone [22] . These prom- of their antibacterial properties.
ising results were confirmed in vivo using a rabbit A remarkable feature making Ga(III) a prom-
scratched cornea infection model: reduction in ising candidate in antibacterial chemotherapy is
P. aeruginosa corneal injury was much more evi- its ability to accumulate within inflammatory
dent upon treatment with DFO–Ga(III)–GEN, sites [4,5,77] . Multiple factors contribute to the
than with individual components [22] . Similarly, accumulation and retention of Ga(III) in inflam-
the β-lactam antibiotic loracarbef, conjugated matory lesions, many of which are attributable to
with DFO–Ga(III), displayed enhanced growth the nonspecific response of injured tissues. Acute
inhibitory effect against Micrococcus luteus, inflammation is characterized by different vas-
compared with individual components [75] . cular events, including vasodilation, increased

future science group www.futuremedicine.com 391


Review  Minandri, Bonchi, Frangipani, Imperi & Visca

permeability and augmented blood flow. Owing in order to improve interlaboratory reproducibil-
to these vascular changes and release of lysoso- ity of results and avoid the bias introduced by
mal content by pathogen-stimulated neutrophils, ill-defined iron content and Ga(III) speciation
high concentration of Tf and Lf are present at in complex bacteriological media.
the site of inflammation [78] . Local production There are a number of open issues that deserve
of bacterial and neutrophil proteases and the low further consideration for the appropriate usage
pH could facilitate Ga(III) release from Tf and of Ga(III) as an antibacterial agent. First of all,
Lf, resulting in remarkably high free Ga(III) con- Ga(III) bioavailability appears to be a major issue,
centration in the infected tissue. These effects, and future research should focus on the devel-
which are the basis of radiogallium scanning opment of optimized Ga(III) carriers endowed
of septic foci [8,9] , could also explain the good with increased capacity to reach infective foci
protective activity of Ga(III) in animal infected and permeate bacterial membrane(s). It should
with bacterial species that show poor sensitivity be taken into account that bacteria generate bio-
to Ga(III) in vitro (Tables 1 & 2) . films in vivo as an adaptive response to the hostile
The wealth of in vitro and in vivo data available environment encountered during the infection.

of
suggest that some Ga(III) formulations could in Bacterial cells embedded in biofilms exist in dif-
the future represent a promising complement ferent metabolic states owing to decreasing oxy-
and/or alternative to the use of conventional gen and nutrient gradients from the surface to the
antimicrobials for treatment of certain bacterial bottom of the biofilm. Cells growing on the bio-

ro
infections. As an example, infections sustained film surface are more metabolically active than
by P. aeruginosa and A. baumannii are extremely those living in deeper layers, and markedly differ
resistant to currently available antibiotic thera- in their response to toxic compounds, including
pies, whilst both species are very sensitive to antimicrobials [79] . Although the biofilm texture
the inhibitory activity of Ga(III) both in vitro contains large water-filled channels that allow
rP
and in vivo. It is tempting to correlate the high free diffusion of ions, the biofilm matrix shows
susceptibility of these species with their obli- a net negative charge due to the presence of
gate respiratory metabolism, where iron plays DNA and anionic polysaccharides that can trap
a pivotal role in electron transfer to and from positively-charged ions such as Ga(III). Thus,
cytochrome complexes. Accordingly, facultative both the resting metabolic state of the bacte-
ho

anaerobes that undertake fermentative processes rial cell and the scavenging effect of the matrix
(e.g., enterobacteria, enterococci, staphylococci could contribute to reducing the antimicrobial
and streptococci) are less susceptible. This poses activity of Ga(III) in biofilm-growing bacteria.
the problem of how the metabolic state of the If translated into clinical practice, higher Ga(III)
cell and the culture condition (e.g., iron levels dosages than those predictable from in vitro anti-
and pO2) affect Ga(III) susceptibility in vitro. microbial susceptibility testing on planktonic
ut

The diverse growth conditions used in vitro to cells would be required for successful therapy of
assess the inhibitory activity of Ga(III) on dif- biofilm-related infections. Related to the above
ferent bacterial pathogens, especially the vari- issues, Ga(III)-coated biomaterials and deliv-
able iron content of culture media, hamper a ery formulations for slow Ga(III) release have
A

straightforward comparison between studies. As recently been developed, providing new tools to
a general rule, the poorer the iron content of the suppress or prevent bacterial growth and biofilm
medium, the higher the Ga(III) activity, irrespec- formation on medical implants [80,81] .
tive of the tested bacterial species. Ga(III) specia- Neither the problem of bacterial resistance to
tion in liquid media should also be taken into Ga(III) nor the duration of bacteriostasis have
account, since extensive Ga(III) precipitation can yet been fully addressed. In vitro activity data for
occur at neutral pH, that is the pH optimum Ga(III) compounds demonstrate bacteriostasis
for growth of bacterial pathogens. Ingredients of for short exposure periods, but it cannot be ruled
culture media can also influence speciation and out that bacteria would outgrow after prolonged
biological effects of Ga(III) by ligand exchange, exposure times, either by mutant selection or
for example between iron complexes and bacte- adaptation. Unlike most antibiotics, which have
rial siderophores [29] . Therefore, it is mandatory a single target within the bacterial cell, Ga(III)
that in vitro antimicrobial susceptibility testing of is a typical multitarget drug that is predicted
Ga(III) compounds are performed in chemically to impair several Fe(III)-dependent functions.
defined media under standard assay conditions, Therefore, resistance mechanisms such as target

392 Future Microbiol. (2014) 9(3) future science group


Promises & failures of gallium as an antibacterial agent  Review

mutation, drug modification or development of eradication. The iron status of the patient should
alternative metabolic pathways are unlikely to also be considered, since pathological iron over-
develop in bacteria to overcome Ga(III) inhibi- load, as in hemochromatosis and thalassemia,
tion. Conversely, reduced permeability by muta- would predictably impair the antibacterial effi-
tion of membrane transporter(s), active efflux, cacy of Ga(III), while chelation therapy has been
extracytoplasmic scavenging (e.g., by sidero- shown to lower Ga(III) levels due to increased
phores) and intracellular trapping into storage urinary secretion [85] . Lastly, nearly all data on
(e.g., bacterioferritin-like) molecules would, the protective activity of Ga(III) in vivo have
in theory, be effective mechanisms for Ga(III) been generated by infecting laboratory animals
resistance; however, for all of them the similarity that had been treated before or concomitant with
between Ga(III) and Fe(III) could imply a severe bacterial challenge [20,23,24,55,60,66,72] and, to our
fitness cost to the resistant cell. This represents knowledge, only one paper has addressed the effi-
another area for future investigation. cacy of Ga(III) administered a few hours postin-
At the molecular level, further studies are fection [26] . These experimental setups are pro-
required to shed more light on the route(s) by foundly different from the clinical setting, where

of
which Ga(III) enters bacterial cells, as well as therapy is usually initiated after the infection is
on the mechanism(s) responsible for growth diagnosed, and call for further in vivo investiga-
inhibition and, eventually, bacterial death. A tions on the efficacy of Ga(III) in the treatment
promising strategy to obtain information on of pre-established bacterial infections.

ro
Ga(III) mechanism(s) of action is the selection of The end of the ‘golden age’ of antibiotic
Ga(III)-insensitive variants upon Ga(III) adap- discovery left us with a dramatic shortage of
tation, and their genetic and functional char- new drugs to combat MDR pathogens [86,87] .
acterization through omics technologies. This Meanwhile, infections caused by MDR bacteria
approach has recently been exploited to address continue to increase in frequency and cause wor-
rP
the mechanisms of Ga(III) resistance in P. aer- ryingly high morbidity and mortality worldwide
uginosa [16] . As to the mechanism of Ga(III)- [88] . In this scenario, the compelling question is
induced bacterial death, it has been shown that whether and when Ga(III) could be used as an
Ga(III) enhances ROS production and generates antibacterial agent. Clinical trials still do not pro-
oxidative stress [11,14] ; however, further research vide the conclusive answer to this question; how-
ho

is needed to determine to what extent these cel- ever, in vitro data and animal infection studies
lular responses play a prominent role in bacte- indicate a potential use of Ga(III) as a last resort
rial death, as proposed for many conventional drug, in combination therapy with other anti-
bactericidal antibiotics [82] . microbials, to treat local or systemic infections
Regarding the clinical applications of Ga(III) sustained by PDR, otherwise untreatable bacte-
as an antibacterial agent, it should be noted ria. The activity of Ga(III) can be potentiated
ut

that Ga(III) is endowed with immunosuppres- in vivo (i.e., the therapeutic dose can be lowered)
sive activity on macrophage and T-cell func- by combination with antibiotics showing syner-
tions [83,84] , and could weaken the host immune gism in vitro (e.g., aminoglycosides and colistin),
response upon long-term administration. The thus expanding the therapeutic potential to those
A

problem is exacerbated in immunocompromised species or strains that are resistant to either of


patients, where the immune deficit impairs effi- the compounds alone. Lastly, the possibility of
cient clearance of infecting bacteria, and who using Ga(III) for topical treatments and coat-
require bactericidal treatments for successful ing of biomaterials and implants is an attractive
eradication of infection. Current therapeutic perspective.
regimens allow attainment of Ga(III) plasma
levels of 10–20 μM, depending on the formula- Financial & competing interests disclosure
tion, with predictably higher levels in infective The authors have no relevant affiliations or financial involve-
foci due to Ga(III) release from Tf and Lf. Such ment with any organization or entity with a financial interest
concentrations are bactericidal for few bacterial in or financial conflict with the subject matter or materials
species and strains (Table 1) , posing the need for discussed in the manuscript. This includes employment, con-
careful determination of the bactericidal concen- sultancies, honoraria, stock ownership or options, expert
tration of Ga(III) compounds against infecting testimony, grants or patents received or pending, or royalties.
bacteria, which should be reached and constantly No writing assistance was utilized in the production of
maintained at the infected site to attain bacterial this manuscript.

future science group www.futuremedicine.com 393


Review  Minandri, Bonchi, Frangipani, Imperi & Visca

Executive summary
Properties of gallium as an antimetabolite
●● Gallium (Ga(III)), is a group IIIA semimetal possessing extensive chemical similarity with iron (Fe(III)).
●● Ga(III) can substitute for Fe(III) in several iron-containing enzymes, host iron-binding proteins, and bacterial
siderophores.
●● Unlike Fe(III), Ga(III) cannot undergo redox cycles under physiological conditions.
●● Replacement of Fe(III) with Ga(III) in the prosthetic group of iron-containing enzymes disrupts essential cellular
functions.
Medical applications of Ga(llI)
●● Ga(III) compounds are used as diagnostic and therapeutic tools in clinical medicine.
●● Citrate-buffered gallium nitrate (Ganite®, Genta, NJ, USA) is approved by US FDA for treatment of cancer-associated
hypercalcemia.

of
Fate of Ga(III) into bacterial cells
●● Ga(III) is likely to exploit Fe(III)-uptake routes to enter bacterial cells.
●● Ga(III) perturbs bacterial iron metabolism and generates oxidative stress.
In vitro & in vivo antibacterial activity of Ga(llI)
●●
ro
Ga(III), in the forms of nitrate, chloride and maltolate salts, is endowed with a potent growth-inhibitory activity, both
in vitro (on planktonic- and biofilm-living bacteria) and in vivo. A number of Gram-negative and Gram-positive bacterial
species, including Acinetobacter baumannii, Pseudomonas aeruginosa, Burkholderia cepacia complex, Mycobacterium
rP
tuberculosis, Francisella tularensis and Rhodococcus equi are severely inhibited by Ga(III).
●● Staphylococcus spp., Salmonella enterica serovars and other facultative anaerobic species are poorly inhibited by Ga(III).
●● Ga(III) activity is influenced by the metabolic state of the cell, and growth inhibition is invariably reversed by Fe(III).
●● A number of Ga(III) complexes have been tested in order to enhance Ga(III) acquisition by bacteria (“Trojan horse”
ho

strategy).
●● Among carriers, Ga(III)–protoporphirin IX showed potent antibacterial activity both in vitro and in vivo.
Considerations for the use of Ga(III) as a new anti-infective
●● Ga(lIl) is more effective against bacterial species that preferentially adopt aerobic metabolism, while facultative
ut

anaerobes undertaking fermentative processes are less susceptible.


●● Pre-existing resistance to conventional antibiotics does not compromise Ga(III) activity.
●● Concerns remain regarding side effects of Ga(III) on host immune response.
A

●● The mechanism(s) of antibacterial action of Ga(III) and its molecular targets deserve more in-depth investigation.
●● Development of a standard method for Ga(III) susceptibility testing will facilitate harmonization of susceptibility data.
●● Definition of minimal bacteriostatic and bactericidal concentrations for individual species and strains is mandatory.

References applications. Nat. Rev. Microbiol. 11(6), 4 Edwards CL, Hayes RL. Tumor scanning
Papers of special note have been highlighted as: 371–384 (2013). with 67Ga citrate. J. Nucl. Med. 10(2),
• of interest •• Outstanding review on the chemical and 103–105 (1969).
•• of considerable interest toxicological principles that underlie the 5 Littenberg RL, Taketa RM, Alazraki NP,
1 Nägeli KW. Über oligodynamische antimicrobial activity of metals. Halpern SE, Ashburn WL. Gallium-67 for
Erscheinungen in lebenden Zellen. Neue localization of septic lesions. Ann. Intern.
3 Adamson RH, Canellos GP, Sieber SM.
Denkschr. Allgemein. Schweiz. Gesellsch. Ges. Med. 79(3), 403–406 (1973).
Studies on the antitumor activity of gallium
Naturweiss. 33, Abt 1 (1893). nitrate (NSC-15200) and other group IIIa 6 Warrell RP, Bockman RS, Coonley CJ, Isaacs
2 Lemire JA, Harrison JJ, Turner RJ. metal salts. Cancer Chemother. Rep. 59(3), M, Staszewski H. Gallium nitrate inhibits
Antimicrobial activity of metals: 599–610 (1975). calcium resorption from bone and is effective
mechanisms, molecular targets and treatment for cancer-related hypercalcemia.
J. Clin. Invest. 73(5), 1487–1490 (1984).

394 Future Microbiol. (2014) 9(3) future science group


Promises & failures of gallium as an antibacterial agent  Review

7 Whitacre C, Apseloff G, Cox K, Matkovic V, 19 Chitambar CR. Medical applications and 28 Peeters E, Nelis HJ, Coenye T. Resistance of
Jewell S, Gerber N. Suppression of toxicities of gallium compounds. Int. J. planktonic and biofilm-grown Burkholderia
experimental autoimmune encephalomyelitis Environ. Res. Public Health. 7(5), 2337–2361 cepacia complex isolates to the transition metal
by gallium nitrate. J. Neuroimmunol. (2010). gallium. J. Antimicrob. Chemother. 61(5),
39(1–2), 175–181 (1992). 20 Antunes LC, Imperi F, Minandri F, Visca P. 1062–1065 (2008).
8 Bernstein LR. Mechanisms of therapeutic In vitro and in vivo antimicrobial activities of 29 Rzhepishevska O, Ekstrand-Hammarström B,
activity for gallium. Pharmacol. Rev. 50(4), gallium nitrate against multidrug-resistant Popp M et al. The antibacterial activity of
665–682 (1998). Acinetobacter baumannii. Antimicrob. Agents Ga 3+ is influenced by ligand complexation as
•• Landmark review on the biochemical, Chemother. 56(11), 5961–5970 (2012). well as the bacterial carbon source. Antimicrob.
• Reports the activity of gallium nitrate in Agents Chemother. 55(12), 5568–5580 (2011).
pharmaceutical and therapeutic properties
of gallium and gallium-based compounds. human serum and in vivo against •• Compares the effects of gallium citrate and
9 Hoffer PB, Huberty J, Khayam-Bashi H. The Acinetobacter baumannii clinical isolates. gallium desferoxamine complexes on the
association of Ga-67 and lactoferrin. J. Nucl. 21 Baldoni D, Steinhuber A, Zimmerli W, growth of several Gram-negative and
Med. 18(7), 713–717 (1977). Trampuz A. In vitro activity of gallium Gram-positive bacterial species.
10 Vallabhajosula SR, Harwig JF, Siemsen JK, maltolate against Staphylococci in logarithmic, 30 Stojiljkovic I, Kumar V, Srinivasan N.
Wolf W. Radiogallium localization in stationary, and biofilm growth phases: Non-iron metalloporphyrins: potent
comparison of conventional and calorimetric antibacterial compounds that exploit haem/

of
tumors: blood binding and transport and the
role of transferrin. J. Nucl. Med. 21(7), susceptibility testing methods. Antimicrob. Hb uptake systems of pathogenic bacteria.
650–656 (1980). Agents Chemother. 54(1), 157–163 (2010). Mol. Microbiol. 31(2), 429–442 (1999).
11 al-Aoukaty A, Appanna VD, Falter H. 22 Banin E, Lozinski A, Brady KM et al. The • Demonstrates strong antibacterial activity of
Gallium toxicity and adaptation in potential of desferrioxamine-gallium as an gallium-protoporphyrin IX against several

ro
Pseudomonas fluorescens. FEMS Microbiol. anti-Pseudomonas therapeutic agent. Proc. Natl Gram-negative and Gram-positive bacterial
Lett. 71(3), 265–272 (1992). Acad. Sci. USA 105(43), 16761–16766 (2008). species.
12 Anghileri LJ, Thuvenot P, Brunotte F, • Describes the in vitro and in vivo activity of 31 Visca P. Iron regulation and siderophore
Marchal C, Robert J. Ionic competition and gallium complexed to the exogenous iron signalling in virulence by Pseudomonas
rP
67
Ga in vivo accumulation. Nuklearmedizin chelator desferoxamine against Pseudomonas aeruginosa. In: Pseudomonas: Virulence and
21(3), 114–116 (1982). aeruginosa. Gene Regulation (Volume 2). Ramos JL (Ed.).
13 Berggren MM, Burns LA, Abraham RT, 23 DeLeon K, Balldin F, Watters C et al. Kluwer Academic Publishers, Dordrecht, The
Powis G. Inhibition of protein tyrosine Gallium maltolate treatment eradicates Netherlands, 69–123 (2004).
phosphatase by the antitumor agent gallium Pseudomonas aeruginosa infection in thermally 32 Faraldo-Gómez JD, Sansom MS. Acquisition
nitrate. Cancer Res. 53(8), 1862–1866 (1993). injured mice. Antimicrob. Agents Chemother. of siderophores in gram-negative bacteria. Nat.
ho

14 Bériault R, Hamel R, Chenier D, Mailloux


53(4), 1331–1337 (2009). Rev. Mol. Cell. Biol. 4(2), 105–116 (2003).
RJ, Joly H, Appanna VD. The overexpression 24 Harrington JR, Martens RJ, Cohen ND, 33 Wandersman C, Delepelaire P. Bacterial iron
of NADPH-producing enzymes counters the Bernstein LR. Antimicrobial activity of sources: from siderophores to hemophores.
oxidative stress evoked by gallium, an iron gallium against virulent Rhodococcus equi Annu. Rev. Microbiol. 58, 611–647 (2004).
mimetic. Biometals 20(2), 165–176 (2007). in vitro and in vivo. J. Vet. Pharmacol. Ther. 34 Cescau S, Cwerman H, Létoffé S, Delepelaire
15 Chitambar CR, Matthaeus WG, Antholine
29(2), 121–127 (2006). P, Wandersman C, Biville F. Heme acquisition
ut

WE, Graff K, O’Brien WJ. Inhibition of 25 Nerren JR, Edrington TS, Bernstein LR et al. by hemophores. Biometals 20(3–4), 603–613
leukemic HL60 cell growth by transferrin- Evaluation of the effect of gallium maltolate (2007).
gallium: effects on ribonucleotide reductase on fecal Salmonella shedding in cattle. J. Food 35 Fabian M, Solomaha E, Olson JS, Maresso
and demonstration of drug synergy with Prot. 74(4), 524–530 (2011). AW. Heme transfer to the bacterial cell
A

hydroxyurea. Blood 72(6), 1930–1936 26 Kaneko Y, Thoendel M, Olakanmi O, envelope occurs via a secreted hemophore in
(1988). Britigan BE, Singh PK. The transition metal the Gram-positive pathogen Bacillus anthracis.
16 García-Contreras R, Lira-Silva E, Jasso- gallium disrupts Pseudomonas aeruginosa iron J. Biol. Chem. 284(46), 32138–32146 (2009).
Chávez R et al. Isolation and characterization metabolism and has antimicrobial and 36 Tiburzi F, Imperi F, Visca P. Is the host heme
of gallium resistant Pseudomonas aeruginosa antibiofilm activity. J. Clin. Invest. 117(4), incorporated in microbial heme-proteins?
mutants. Int. J. Med. Microbiol. 303(8), 877–888 (2007). IUBMB Life 61(1), 80–83 (2009).
574–582 (2013). •• Seminal work on the activity of gallium 37 Schalk IJ, Guillon L. Fate of ferrisiderophores
17 Hedley DW, Tripp EH, Slowiaczek P, Mann nitrate in vitro and in vivo against after import across bacterial outer membranes:
GJ. Effect of gallium on DNA synthesis by P. aeruginosa clinical isolates. different iron release strategies are observed in
human T-cell lymphoblasts. Cancer Res. the cytoplasm or periplasm depending on the
27 Olakanmi O, Britigan BE, Schlesinger LS.
48(11), 3014–3018 (1988). siderophore pathways. Amino Acids 44(5),
Gallium disrupts iron metabolism of
18 Bernstein LR. 31Ga therapeutic gallium mycobacteria residing within human 1267–1277 (2013).
compounds. In: Metallotherapeutic Drugs and macrophages. Infect. Immun. 68(10), 38 Cartron ML, Maddocks S, Gillingham P,
Metal-Based Diagnostic Agents: the Use of 5619–5627 (2000). Craven CJ, Andrews SC. Feo-transport of
Metals in Medicine. Gielen M, Tiekink ERT ferrous iron into bacteria. Biometals 19(2),
▪ Relevant publication on the effect gallium
(Eds). John Wiley & Sons, Ltd, Chichester, 143–157 (2006).
nitrate on intracellular mycobacteria.
UK, 259–277 (2005).

future science group www.futuremedicine.com 395


Review  Minandri, Bonchi, Frangipani, Imperi & Visca

39 Andrews SC, Robinson AK, Rodríguez- catechol-substituted cephalosporin is Emerging therapies for multidrug resistant
Quiñones F. Bacterial iron homeostasis. unrelated to the pyochelin-Fe transporter Acinetobacter baumannii. Trends Microbiol.
FEMS Microbiol. Rev. 27(2–3), 215–237 FptA. Amino Acids. 38(5), 1627–1629 (2010). 21(3), 157–163 (2013).
(2003). 51 Imperi F, Massai F, Facchini M et al. 62 Fecteau ME, Fyock TL, McAdams SC,
40 Oglesby-Sherrouse AG, Murphy ER. Repurposing the antimycotic drug flucytosine Boston RC, Whitlock RH, Sweeney RW.
Iron-responsive bacterial small RNAs: for suppression of Pseudomonas aeruginosa Evaluation of the in vitro activity of gallium
variations on a theme. Metallomics 5(4), pathogenicity. Proc. Natl Acad. Sci. USA nitrate against Mycobacterium avium subsp
276–286 (2013). 110(18), 7458–7463 (2013). paratuberculosis. Am. J. Vet. Res. 72(9),
41 Salvail H, Massé E. Regulating iron storage 52 Braun V. Iron uptake by Escherichia coli. 1243–1246 (2011).
and metabolism with RNA: an overview of Front. Biosci. 8, 1409–1421 (2003). 63 Bernstein LR, Tanner T, Godfrey C, Noll B.
posttranscriptional controls of intracellular 53 Braud A, Hannauer M, Mislin GL, Schalk IJ. Chemistry and pharmacokinetics of gallium
iron homeostasis. Wiley Interdiscip. Rev. RNA The Pseudomonas aeruginosa pyochelin-iron maltolate, a compound with high oral
3(1), 26–36 (2012). uptake pathway and its metal specificity. J. gallium bioavailability. Met. Based Drugs
42 Litwin CM, Calderwood SB. Role of iron in Bacteriol. 191(11), 3517–3525 (2009). 7(1), 33–47 (2000).
regulation of virulence genes. Clin. Microbiol. 54 Braud A, Hoegy F, Jezequel K, Lebeau T, 64 Arnold CE, Bordin A, Lawhon SD, Libal
Rev. 6(2), 137–149 (1993). Schalk IJ. New insights into the metal MC, Bernstein LR, Cohen ND.
43 García-Contreras R, Pérez-Eretza B, Lira-Silva specificity of the Pseudomonas aeruginosa Antimicrobial activity of gallium maltolate

of
E et al. Gallium induces the production of pyoverdine-iron uptake pathway. Environ. against Staphylococcus aureus and methicillin-
virulence factors in Pseudomonas aeruginosa. Microbiol. 11(5), 1079–1091 (2009). resistant S. aureus and Staphylococcus
Pathog. Dis. (doi:10.1111/2049–632X.12105. pseudintermedius: an in vitro study. Vet.
55 Bozja JYK, Shafer WM, Stojiljkovic I.
(2013) (Epub ahead of print). Microbiol. 155(2–4), 389–394 (2012).
Porphyrin-based compounds exert

ro
44 Krueger S, Olson GJ, Johnsonbaugh D, antibacterial action against the sexually 65 Miranda-Casoluengo R, Coulson GB,
Beveridge TJ. Characterization of the binding transmitted pathogens Neisseria gonorrhoeae Miranda-Casoluengo A, Vázquez-Boland JA,
of gallium, platinum, and uranium to and Haemophilus ducreyi. Int. J. Antimicrob. Hondalus MK, Meijer WG. The
Pseudomonas fluorescens by small-angle X-ray Agents. 24(6), 578–584 (2004). hydroxamate siderophore rhequichelin is
scattering and transmission electron required for virulence of the pathogenic
rP
56 Olczak T, Maszczak-Seneczko D, Smalley
microscopy. Appl. Environ. Microbiol. 59(12), actinomycete Rhodococcus equi. Infect.
JW, Olczak M. Gallium(III), cobalt(III) and
4056–4064 (1993). Immun. 80(12), 4106–4114 (2012).
copper(II) protoporphyrin IX exhibit
45 Anderson DS, Adhikari P, Nowalk AJ, Chen antimicrobial activity against Porphyromonas 66 Olakanmi O, Gunn JS, Su S, Soni S, Hassett
CY, Mietzner TA. The hFbpABC transporter gingivalis by reducing planktonic and biofilm DJ. Gallium disrupts iron uptake by
from Haemophilus influenzae functions as a growth and invasion of host epithelial cells. intracellular and extracellular Francisella
binding-protein-dependent ABC transporter Arch. Microbiol. 194(8), 719–724 (2012). strains and exhibits therapeutic efficacy in a
ho

with high specificity and affinity for ferric murine pulmonary infection model.
57 Wójtowicz H, Bielecki M, Wojaczyski J,
iron. J. Bacteriol. 186(18), 6220–6229 Antimicrob. Agents Chemother. 54(1),
Olczak M, Smalley JW, Olczak T. The
(2004). 244–253 (2010).
Porphyromonas gingivalis HmuY haemophore
46 Foley TL, Simeonov A. Targeting iron binds gallium(III), zinc(II), cobalt(III), 67 Pan X, Tamilselvam B, Hansen EJ, Daefler S.
assimilation to develop new antibacterials. manganese(III), nickel(II), and copper(II) Modulation of iron homeostasis in
Expert Opin. Drug Discov. 7(9), 831–847 protoporphyrin IX but in a manner different macrophages by bacterial intracellular
ut

(2012). to iron(III) protoporphyrin IX. Metallomics pathogens. BMC Microbiol. 10, 64 (2010).

47 Kontoghiorghes GJ, Kolnagou A, Skiada A, 5(4), 343–351 (2013). 68 Theurl I, Fritsche G, Ludwiczek S, Garimorth
Petrikkos G. The role of iron and chelators on 58 Stojiljkovic I, Evavold BD, Kumar V. K, Bellmann-Weiler R, Weiss G. The
infections in iron overload and non-iron Antimicrobial properties of porphyrins. macrophage: a cellular factory at the
A

loaded conditions: prospects for the design of Expert Opin. Investig. Drugs. 10(2), 309–320 interphase between iron and immunity for the
new antimicrobial therapies. Hemoglobin (2001). control of infections. Biometals 18(4),
34(3), 227–239 (2010). 359–367 (2005).
59 Magiorakos AP, Srinivasan A, Carey RB
48 Visca P, Bonchi C, Minandri F, Frangipani E, et al. Multidrug-resistant, extensively 69 Coleman M, Kuskie K, Liu M et al. In vitro
Imperi F. The dual personality of iron drug-resistant and pandrug-resistant bacteria: antimicrobial activity of gallium maltolate
chelators: growth inhibitors or promoters? an international expert proposal for interim against virulent Rhodococcus equi. Vet.
Antimicrob. Agents Chemother. 57(5), standard definitions for acquired resistance. Microbiol. 146(1–2), 175–178 (2010).
2432–2433 (2013). Clin. Microbiol. Infect. 18(3), 268–281 70 Martens RJ, Mealey K, Cohen ND et al.
49 Kelson AB, Carnevali M, Truong-Le V. (2012). Pharmacokinetics of gallium maltolate after
Gallium-based anti-infectives: targeting 60 de Léséleuc L, Harris G, KuoLee R, Chen W. intragastric administration in neonatal foals.
microbial iron-uptake mechanisms. Curr. In vitro and in vivo biological activities of Am. J. Vet. Res. 68(10), 1041–1044 (2007).
Opin. Pharmacol. 13(5), 707–771 (2013). iron chelators and gallium nitrate against 71 Levaditi C, Bardet J, Tchakirian A, Vaisman

• Outlines the feasibility to use gallium-based Acinetobacter baumannii. Antimicrob. Agents A. Le gallium, propriétés thérapeutiques dans
Chemother. 56(10), 5397–5400 (2012). la syphilis et les trypanosomiases
complexes against bacteria.
61 García-Quintanilla M, Pulido MR, expérimentales. C. R. Hebd. Seances Acad. Sci.
50 Hoegy F, Gwynn MN, Schalk IJ. Ser. D. Sci. Nat. 192, 1142–1143 (1931).
López-Rojas R, Pachón J, McConnell MJ.
Susceptibility of Pseudomonas aeruginosa to

396 Future Microbiol. (2014) 9(3) future science group


Promises & failures of gallium as an antibacterial agent  Review

72 Fecteau ME, Whitlock RH, Fyock TL, 77 Lavender JP, Lowe J, Barker JR, Burn JI, in vivo T-cell responses by transferrin-gallium
McAdams SC, Boston RC, Sweeney RW. Chaudhri MA. Gallium 67 citrate scanning and gallium nitrate. Blood 88(8), 3056–3064
Antimicrobial activity of gallium nitrate in neoplastic and inflammatory lesions. Br. J. (1996).
against Mycobacterium avium subsp. Radiol. 44(521), 361–366 (1971). 84 Huang EH, Gabler DM, Krecic ME, Gerber
paratuberculosis in neonatal calves. J. Vet. 78 Tsan MF. Studies on gallium accumulation in N, Ferguson RM, Orosz CG. Differential
Intern. Med. 25(5), 1152–1155 (2011). inflammatory lesions: III. Roles of effects of gallium nitrate on T lymphocyte
73 Goss CH, Hornick DB, Aitken ML, polymorphonuclear leukocytes and bacteria. and endothelial cell activation.
Anderson G, Caldwell E. Phase 1 J. NucI. Med. 19(5), 492–495 (1978). Transplantation 58(11), 1216–1222 (1994).
pharmacokinetic and safety study of 79 de la Fuente-Núñez C, Reffuveille F, 85 Nagamachi S, Hoshi H, Jinnouchi S, Ono S,
intravenous GaniteTM (Gallium nitrate) in CF. Fernández L, Hancock RE. Bacterial biofilm Watanabe K. Gallium-67 scintigraphy in
Pediatric Pulmunol. 47(Suppl. 35), 303 development as a multicellular adaptation: patients with hemochromatosis treated by
(2012). antibiotic resistance and new therapeutic deferoxamine. Ann. Nucl. Med. 2(1), 35–39
74 Principe L, D’Arezzo S, Capone A, Petrosillo strategies. Curr. Opin. Microbiol. 16(5), (1988).
N, Visca P. In vitro activity of tigecycline in 580–589 (2013). 86 Boucher HW, Talbot GH, Benjamin DK
combination with various antimicrobials 80 Valappil SP, Ready D, Abou Neel EA et al. et al. Infectious Diseases Society of America.
against multidrug resistant Acinetobacter Antimicrobial gallium-doped phosphate- 10x’20 Progress-development of new drugs
baumannii. Ann. Clin. Microbiol. Antimicrob. based glasses. Adv. Funct. Mater. 18(5), active against gram-negative bacilli: an update

of
21(8), 18 (2009). 732–741 (2008). from the Infectious Diseases Society of
75 Juárez-Hernández RE, Miller PA, Miller MJ. 81 Valappil SP, Yiu HH, Bouffier L et al. Effect America. Clin. Infect. Dis. 56(12), 1685–1694
Syntheses of siderophore-drug conjugates of novel antibacterial gallium-carboxymethyl (2013).
using a convergent thiol-maleimide system. cellulose on Pseudomonas aeruginosa. Dalton 87 Theuretzbacher U. Accelerating resistance,

ro
ACS Med. Chem. Lett. 11(10), 799–803 Trans. 42(5), 1778–1786 (2013). inadequate antibacterial drug pipelines and
(2012). international responses. Int. J. Antimicrob.
82 Kohanski MA, Dwyer DJ, Hayete B,
76 Halwani M, Yebio B, Suntres ZE, Alipour M, Lawrence CA, Collins JJ. A common Agents 39(4), 295–299 (2012).
Azghani AO, Omri A. Co-encapsulation of mechanism of cellular death induced by 88 Boucher HW, Talbot GH, Bradley JS et al.
gallium with gentamicin in liposomes bactericidal antibiotics. Cell 130(5), 797–810 Bad bugs, no drugs: no ESKAPE! An update
rP
enhances antimicrobial activity of gentamicin (2007). from the Infectious Diseases Society of
against Pseudomonas aeruginosa. J. Antimicrob. America. Clin. Infect. Dis. 48(1), 1–12
83 Drobyski WR, Ul-Haq R, Majewski D,
Chemother. 62(6), 1291–1297 (2008). (2009).
Chitambar CR. Modulation of in vitro and
ho
ut
A

future science group www.futuremedicine.com 397

View publication stats

You might also like