Skip to main content
Infection and Immunity logoLink to Infection and Immunity
. 2010 Feb 16;78(5):2045–2052. doi: 10.1128/IAI.01236-09

The Yersiniabactin Transport System Is Critical for the Pathogenesis of Bubonic and Pneumonic Plague

Jacqueline D Fetherston 1, Olga Kirillina 1, Alexander G Bobrov 1, James T Paulley 1, Robert D Perry 1,*
PMCID: PMC2863531  PMID: 20160020

Abstract

Iron acquisition from the host is an important step in the pathogenic process. While Yersinia pestis has multiple iron transporters, the yersiniabactin (Ybt) siderophore-dependent system plays a major role in iron acquisition in vitro and in vivo. In this study, we determined that the Ybt system is required for the use of iron bound by transferrin and lactoferrin and examined the importance of the Ybt system for virulence in mouse models of bubonic and pneumonic plague. Y. pestis mutants unable to either transport Ybt or synthesize the siderophore were both essentially avirulent via subcutaneous injection (bubonic plague model). Surprisingly, via intranasal instillation (pneumonic plague model), we saw a difference in the virulence of Ybt biosynthetic and transport mutants. Ybt biosynthetic mutants displayed an ∼24-fold-higher 50% lethal dose (LD50) than transport mutants. In contrast, under iron-restricted conditions in vitro, a Ybt transport mutant had a more severe growth defect than the Ybt biosynthetic mutant. Finally, a Δpgm mutant had a greater loss of virulence than the Ybt biosynthetic mutant, indicating that the 102-kb pgm locus encodes a virulence factor, in addition to Ybt, that plays a role in the pathogenesis of pneumonic plague.


Nearly all organisms require trace amounts of iron. Pathogens must overcome host iron- and heme-binding proteins to cause an infection and disease. The importance of iron acquisition mechanisms has been demonstrated in a number of bacterial pathogens (14, 15, 27, 32, 85). Yersinia pestis, the causative agent of plague, has a number of proven and putative iron and heme transport systems. Of these systems, the yersiniabactin (Ybt) siderophore-dependent iron transport system plays a major role in the virulence of bubonic plague in mice (7, 8, 38, 76, 79).

All identified genes required for the regulation, synthesis, and transport of Ybt, except for ybtD, are carried within a high-pathogenicity island (HPI) that has been spread among enteric pathogens but is essentially identical in the pathogenic yersiniae (13, 59, 79). In Y. pestis, the ∼36-kb HPI is located within the 102-kb pgm locus; the entire pgm locus undergoes spontaneous deletion in vitro at a frequency of about 10−5 (25, 39, 62). The Ybt system produces a siderophore composed of one salicylate, one thiazoline, and two thiazolidine rings via a nonribosomal peptide/polyketide synthesis mechanism involving high-molecular-weight protein 1 (HMWP1), HMWP2, YbtD, YbtT, YbtE, YbtU, and YbtS (76, 79, 94). The formation constant of this siderophore with ferric iron is 4 × 1036, and the crystal structure of the ferric complex has been solved (68, 78).

Iron from the Ybt-Fe complex is transported into the cell via the TonB-dependent outer membrane (OM) receptor Psn (which is also required for sensitivity to the bacteriocin pesticin) and an ABC transporter consisting of two inner membrane (IM), fused-function permease/ATP-binding proteins, YbtP and YbtQ. A mutation in any of these three genes prevents Ybt-dependent uptake of iron but does not prevent Ybt secretion. YbtX is encoded in an apparent four-gene operon (ybtPQXS) and is a predicted IM protein which resembles an exporter, with 12 predicted transmembrane domains (37, 76, 79, 80). It has weak homology to Escherichia coli EntS and Bordetella AlcS, exporters for enterobactin and alcaligin, respectively (18, 44, 83), but stronger similarities to RhtX and FptX, which import rhizobactin and pyochelin in Sinorhizobium meliloti and Pseudomonas aeruginosa, respectively (66, 70). A mutation in ybtX does not cause a significant defect in either Ybt synthesis or the ability to use Ybt as an iron source. Thus, the role of YbtX, if any, in the Ybt system remains an enigma (7, 37, 38, 76, 79).

In addition to typical Fur-Fe repression, maximal activation of the Ybt biosynthetic and transport operons requires an AraC-like positive regulator, YbtA, and its cognate siderophore, Ybt. Similar mechanisms activate diverse siderophore systems in a number of bacteria (21, 36, 47, 63, 65, 69, 76, 77, 79).

Here we show that the Ybt system can remove iron from transferrin and lactoferrin. In addition we examine the role of the Ybt system in the pathogenesis of plague in mice. Previously we showed that the Ybt system was required for virulence by a subcutaneous (s.c.) route of infection using an attenuated strain of Y. pestis (yopJ psa). Strains bearing mutations in the OM receptor (psn), the IM permease/ATPase (ybtP), or a Ybt biosynthetic enzyme (irp2) failed to kill mice at the highest doses tested (7, 37). We have obtained similar results in this study with a fully virulent strain. In addition, our experiments indicate that the Ybt system is also important in pneumonic plague. However, there were interesting differences in the virulence of siderophore receptor (psn) mutants and biosynthetic (irp2) mutants which were not seen in s.c. infections. Strains which produce the siderophore but are unable to use it (i.e., psn mutants) were more virulent than the biosynthetic mutants. However, in vitro the psn mutant was more defective than the biosynthetic mutant for growth under iron-restricted conditions.

MATERIALS AND METHODS

Bacterial strains and cultivation.

The bacterial strains and plasmids used in this study are listed in Table 1. From glycerol stocks (10), Y. pestis strains were grown on Congo red (CR) agar (88) before being transferred to tryptose blood agar base (TBA) slants. Formation of red colonies on CR plates indicates that the strain has retained the pgm locus, which can be spontaneously lost at a rate of 10−5 (25, 39, 62).

TABLE 1.

Y. pestis strains and plasmids used in this study

Strain or plasmid Relevant characteristics Reference or source
Strainsa
    KIM5 Pgmpgm; Hms Ybt) Lcr+ Pla+; pMT1, pCD1, pPCP1 92
    KIM5(pCD1Ap)+ Apr Pgm+ (Hms+ Ybt+) Lcr+ Pla+; pMT1, pCD1Ap (′yadA::bla), pPCP1; derived from KIM6+ 45
    KIM5(pCD1Ap) Apr Pgmpgm; Hms Ybt) Lcr+ Pla+; pMT1, pCD1Ap (′yadA::bla), pPCP1; derived from KIM6 This study
    KIM5-2045.1 (pCD1Ap) Apr Hms+ Ybtpsn2045.1) Lcr+ Pla+; pMT1, pCD1Ap (′yadA::bla), pPCP1; derived from KIM6-2045.1 This study
    KIM5-2045.6 (pCD1Ap) Kmr Apr Hms+ Ybtpsn::kan2045.6) Lcr+ Pla+; pMT1, pCD1Ap (′yadA::bla), pPCP1; derived from KIM6-2045.6 This study
    KIM5-2046.1 (pCD1Ap) Kmr Apr Hms+ Ybt (irp2::kan2046.1) Lcr+ Pla+; pMT1, pCD1Ap (′yadA::bla), pPCP1; derived from KIM6-2046.1 This study
    KIM5-2046.3 (pCD1Ap) Apr Hms+ Ybtirp2-2046.3) Lcr+ Pla+; pMT1, pCD1Ap (′yadA::bla), pPCP1; derived from KIM6-2046.3 This study
    KIM6+ Pgm+ (Hms+ Ybt+) Lcr Pla+; pMT1, pPCP1 38
    KIM6 Pgmpgm; Hms Ybt) Lcr Pla+; pMT1, pPCP1 38
    KIM6-2045.1 Hms+ Ybtpsn2045.1) Lcr Pla+; pMT1, pPCP1 38
    KIM6-2045.6 Kmr Hms+ Ybtpsn::kan2045.6) Lcr Pla+; pMT1, pPCP1 36
    KIM6-2046.1 Kmr Hms+ Ybt (irp2::kan2046.1) Lcr Pla+; pMT1, pPCP1 38
    KIM6-2046.3 Hms+ Ybtirp2-2046.3) Lcr Pla+; pMT1, pPCP1 7
    KIM6-2180 Hms+ Ybtirp2-2046.3 Δpsn2045.1) Lcr Pla+; pMT1, pPCP1 This study
Plasmids
    pCD1Ap 71.7 kb, Apr Lcr+; pCD1 with bla cassette inserted into ′yadA downstream of the frameshift mutation in this pseudogene 45
    pCSIRP498.9 8.7 kb, Apr Sucs (sacB+), Δirp2-2046.3, R6K origin suicide vector 7
a

A plus sign indicates an intact chromosomal 102-kb pgm locus. All other Y. pestis strains have a mutation within this locus or a deletion of the entire locus.

For iron-deficient growth studies, Y. pestis cells were harvested from TBA slants and grown in chemically defined medium (PMH or PMH2) which had been extracted prior to use with Chelex 100 resin (Bio-Rad Laboratories). A previously published paper by Gong et al. has an error in the published buffer concentrations; the concentrations of PIPES [piperazine-N,N′-bis(2-ethanesulfonic acid)] and HEPES should be 50 mM for PMH2 and PMH, respectively (45, 86). For iron-replete growth, Y. pestis strains were cultivated in PMH or PMH2 supplemented with 10 μM FeCl3. Growth of the cultures was monitored by determining the optical density at 620 nm (OD620) with a Genesys5 spectrophotometer (Spectronic Instruments, Inc.). Growth through two transfers (∼6 to 8 generations) was used to acclimate cells to PMH2 and varying iron conditions prior to use in experimental studies.

All glassware used for iron-restricted studies was soaked overnight in ScotClean (OWL Scientific, Inc.) to remove contaminating iron and copiously rinsed in deionized water. Where appropriate, ampicillin (Ap) (50 to 100 μg/ml) or kanamycin (Km) (50 μg/ml) was added to media.

Construction of Y. pestis KIM6-2180 (Δpsn2045.1 Δirp2-2046.3).

Suicide plasmid pCSIRP498.9 (encoding an in-frame Δirp2-2046.3 mutation) was electroporated into Y. pestis KIM6-2045.1 (Δpsn2045.1), and merodiploids were selected on TBA plates containing Ap. As previously described (8), selected cointegrants were grown overnight in heart infusion broth (HIB) without Ap and plated on CR plates containing 5% sucrose to select for recombinants with the Δirp2-2046.3 mutation. The mutation was confirmed by Southern blot analysis (data not shown), and the Δirp2-2046.3 Δpsn2045.1 double mutant was designated Y. pestis strain KIM6-2180 (Table 1).

Plasmids and DNA techniques.

Plasmids were purified by alkaline lysis from cultures grown overnight in HIB (12). Y. pestis cells were transformed by electroporation as previously described (38).

Assay for use of Tf and Lf.

Y. pestis strains KIM6+ (Pgm+) and KIM6-2046.1 (irp2::kan2046.1) were grown through two transfers at 37°C in deferrated PMH with or without 0.5 mM NaHCO3 (PMH-NaHCO3). Second-transfer overnight cultures were used to seed 20 ml of molten PMH-NaHCO3-1% agarose with ∼5 × 106 cells, and ethylenediamine-di(o-hydroxyphenyl-acetic acid) (EDDA) was added to a final concentration of 15 μM or 7.5 μM to inhibit the growth of KIM6+ or KIM6-2046.1 (irp2::kan2046.1), respectively, on plates for transferrin (Tf) growth responses. For tests of lactoferrin (Lf) growth responses, 50 μM EDDA was used to inhibit the growth of KIM6+ and KIM6-2046.3 (Δirp2-2046.3). Aliquots of 15 μl of various iron sources were placed on the plates: partially iron-saturated Tf (Sigma; 50 mg/ml) or 1 mM FeCl3 was added to 1-mm-diameter wells in the agar, while bovine Lf (Sigma; 50 mg/ml) or 1 mM FeSO4 was placed on filter discs. Alternatively, PMH-NaHCO3-EDDA plates seeded with Y. pestis strains were overlaid with a dialysis membrane (12,000- to 14,000-Da molecular mass cutoff), and mixtures of PMH, 2% agarose, and the above-described iron-containing preparations were placed as a drop on the dialysis membrane. Although no information on the degree of iron saturation of bovine Lf is provided by Sigma, it is ∼15 to 20% saturated in its natural state (87). To remove free iron, Lf resuspended in a dialysis buffer (0.1 M sodium citrate-0.1 M sodium bicarbonate, pH 7.2) with 0.4% sodium azide at 150 mg/ml was dialyzed (12,000- to 14,000-Da molecular mass cutoff) against 500 ml of this buffer containing 0.4% sodium azide and 20 μM EDDA for 40 min with two buffer changes. Following overnight dialysis at 4°C in the buffer without EDDA or sodium azide, the Lf solution was adjusted to 50 mg/ml in PMH2 containing 2 mM sodium bicarbonate. Bacterial growth on the plates was monitored over 3 days for Tf and 2 days for Lf and visualized by overlayering the plate with TBA containing 2 mM ferric citrate and 1.5 mM esculin. Y. pestis hydrolyzes esculin to produce a black precipitate in areas of growth.

Urea gel electrophoresis of iron-Tf complexes.

Y. pestis strains KIM6+ and KIM6-2046.1 were grown at 37°C through three transfers in iron-depleted PMH-NaHCO3. Dialysis baggies (12,000- to 14,000-Da molecular mass cutoff) containing iron-saturated Tf at a concentration of 1 mg/ml in PMH2 were placed in a suspension of Y. pestis cells in PMH-NaHCO3 or in uninoculated medium as a negative control. After overnight incubation at 37°C, a 10-μl aliquot of the Tf samples was mixed with sample buffer and electrophoresed through a 6.5 M urea-6% polyacrylamide gel in Tris-borate buffer (pH 8.3) as previously described (33, 64). To identify changes in transferrin saturation, partially saturated Tf (a mixture of N- and C-end iron-loaded monosaturated forms) (Sigma) was electrophoresed, as well as an equimolar mixture of holo-Tf (fully iron saturated) (ICN Biomedicals Inc.). The proteins were visualized using Coomassie blue staining.

Virulence testing.

Construction of potentially virulent strains and virulence testing were performed in a CDC-approved biosafety level 3 (BSL3) laboratory following select agent regulations, using procedures approved by the University of Kentucky Institutional Biosafety Committee. Y. pestis strains were transformed with the virulence plasmid pCD1Ap by electroporation (43, 45) and plated on TBA plates containing Ap (50 μg/ml). The plasmid profile of transformants was analyzed, as well as their phenotype on CR agar (88) and magnesium-oxalate plates (48). Supernatants from cultures grown at 37°C in the absence of CaCl2 were tested for the secretion of LcrV by Western blot analysis using polyclonal antisera against histidine-tagged LcrV (40, 43).

For s.c. infections, overnight cultures of Y. pestis cells grown in HIB at 26°C were diluted to an OD of 0.1 at 620 nm and incubated in HIB at 26°C until they reached mid-logarithmic phase (OD of ∼0.5). Samples were harvested and diluted in mouse isotonic PBS (149 mM NaCl, 16 mM Na2HPO4, 4 mM NaH2PO4). Groups of four 6- to 8-week-old female Swiss Webster (Hsd::ND4) mice were injected subcutaneously with 0.1 ml of 10-fold serially diluted bacterial suspensions ranging from 100 to 108 CFU/ml.

Cells used for intranasal (i.n.) instillation infections were grown at 37°C in HIB containing 4 mM CaCl2 to prevent full induction of Lcr in vitro and were similarly diluted in mouse isotonic PBS. Twenty microliters of the bacterial suspension was administered to the nares (∼5-μl doses alternating between the two nostrils) of mice sedated with 100 μg of ketamine and 10 μg of xylazine per kg of body weight. The actual i.n. and s.c. bacterial doses administered were determined by plating aliquots of serially diluted suspensions of each dose, in duplicate, onto TBA plates containing Ap (50 μg/ml). The colonies were counted on plates incubated at 30°C for 2 days. Mice were observed daily for 2 weeks, and 50% lethal doses (LD50s) were calculated according to the method of Reed and Muench (84). All animal care and experimental procedures were conducted in accordance with the Animal Welfare Act, the Guide for the Care and Use of Laboratory Animals (69b), the Public Health Service Policy on Humane Care and Use of Laboratory Animals (69a), and the U.S. Government Principles for the Utilization and Care of Vertebrate Animals Used in Teaching, Research, and Training (70a) and were approved by the University of Kentucky Institutional Animal Care and Use Committee. The University of Kentucky Animal Care Program first obtained accreditation from the Association for the Assessment and Accreditation of Laboratory Animal Care, Inc. (AAALAC), in 1966 and has maintained full accreditation continuously since that time.

RESULTS

Ybt-dependent use of host iron-binding proteins.

We tested iron-stressed cells of Y. pestis KIM6+ (Ybt+), KIM6-2046.1 (irp2::kan2046.1; Ybt), or KIM6-2046.3 (in-frame Δirp2-2046.3; Ybt) for their ability to use Tf and Lf as sole sources of iron for growth on PMH-NaHCO3-EDDA plates. A functional Ybt system allowed the use of both host proteins and inorganic iron even when the cells were separated from the solutions by a dialysis membrane. Inorganic iron stimulated the growth of the Ybt mutant. The Ybt mutant was unable to use either Tf or Lf as an iron source (Fig. 1).

FIG. 1.

FIG. 1.

Use of iron from transferrin (Tf) and lactoferrin (Lf). Growth responses of Ybt+ KIM6+ (A and C), Ybt KIM6-2046.1 (irp2::kan2046.1) (B), and KIM6-2046.3 (Δirp2-2046.3) (D) on PMH-EDDA plates to partially saturated Tf (1, 2, 5, and 6 in panels A and B), partially saturated bovine Lf (1, 2, 5, and 6 in panels C and D), or inorganic iron (3, 4, 7, and 8) are shown. The solutions were added to wells (A and B) or on filter discs (C and D) on seeded plates (1, 3, 5, and 7) or spotted onto a dialysis membrane overlaying the bacterial cells (2, 4, 6, and 8). After incubation with Tf or Lf at 37°C, plates were overlaid with TBA containing esculin and ferric citrate to visualize bacterial growth. The images are from one of two or more independent experiments that yielded similar results.

We also demonstrated the ability of Ybt to remove iron from Tf. The mobility of Tf in polyacrylamide gels containing urea is affected by iron saturation and which of the two iron-binding sites are filled (33, 41, 50, 64, 96). Iron-saturated Tf was separated from cultures by a dialysis baggy and incubated overnight at 37°C. Subsequent urea gel electrophoresis of the Tf solution showed that incubation with a KIM6+ culture converted the majority of Fe-saturated Tf to a less saturated form (Fig. 2, lane 2). In contrast, the Y. pestis irp2 mutant was able to convert only a small amount of saturated Tf to an Fe-Tf form; this conversion was not seen in uninoculated medium (Fig. 2, compare lanes 1 lane 3). Nevertheless, these results indicate that Ybt is able to directly remove iron from Tf as well as use iron from Lf.

FIG. 2.

FIG. 2.

Removal of iron from iron-saturated human transferrin by Y. pestis. Tf was incubated overnight with iron-starved cultures of Y. pestis Ybt+ KIM6+ (lane 2) or Ybt KIM6-2046.1 (lane 3). Tf was in a dialysis baggy (12,000- to 14,000-Da molecular mass cutoff), separating it from the cultures. After incubation, the various iron-saturated Tf forms were separated on 6% polyacrylamide-urea gels. Incubation of Tf with PMH growth medium (lane 1) was used as a negative control. Partially saturated Tf (lane 4) and a mixture of various forms of Tf (lane 5) were used to show mobility shifts due to iron saturation of Tf. The image is from one of two independent experiments, which yielded similar results.

Ybt and bubonic plague.

Previously we tested the virulence of various iron transport mutants using mildly attenuated strains of Y. pestis with mutations in yopJ and psa. In this study we used a reconstructed wild-type (WT) strain [KIM5(pCD1Ap)+] to test the effects of Ybt transport and biosynthesis mutations on virulence. Table 2 shows the LD50s in mice infected by an s.c. route. Twenty-five cells killed 50% of mice infected with this WT strain; with doses greater than the LD50, mice had ruffled fur starting on day 3 postinfection and developed a hunched posture by day 5, with deaths occurring between days 5 and 13. In contrast, doses of ∼107 for both the Ybt transport and biosynthesis mutants caused transient illness (ruffled fur and hunched posture at the highest doses), but only 1 of 16 mice infected with the Δirp2 biosynthetic mutant died (day 5 postinfection). Note that two independent mutations in psn and irp2 yielded similar results, indicating that the loss of virulence is not due to an unidentified secondary mutation in these strains. Consequently, the Ybt system is critical for virulence by this route of infection, with mutants displaying a >4.3 × 105-fold loss of virulence in this model.

TABLE 2.

LD50s for Y. pestis strains in mouse models of pneumonic and bubonic plague

Model Strain or mutation LD50 (mean ± SD)a
Pneumonic plague Wild type 329 ± 105
Δpsn2045.1 or psn::kan2045.6 1.1 × 104 ± 2.9 × 103
Δirp2-2046.3 or irp2::kan2046.1 2.6 × 105 ± 1.8 × 105
Δpgmb >3.9 × 106
Bubonic plague Wild type 25 ± 12
Δpsn2045.1 or psn::kan2045.6 >2.6 × 107
Δirp2-2046.3 or irp2::kan2046.1 >1.3 × 107
a

> indicates an LD50 above the highest bacterial doses tested from at least two independent experiments. Values were calculated from two or more independent trials. Probit analysis using SPSS determined that the intranasal LD50s of the three mutants are significantly different from each other (see text). For statistical analysis, the Δpgm LD50 was set at 3.9 × 106.

b

Two Δpgm strains [KIM5 and KIM5(pCD1Ap) (Table 1)] with a native and recombinant pCD1 were tested but showed the same loss of virulence.

Ybt and pneumonic plague.

An i.n. route has been used to initiate pneumonic infections by Brucella melitensis, Chlamydia trachomatis, Francisella tularensis, P. aeruginosa, Streptococcus pneumoniae, and Y. pestis (3, 4, 28, 32, 51, 54, 56, 72, 81, 90, 97). Latham et al. (54) noted that the i.n. model in mice causes a severe bronchopneumonia that closely resembles descriptions of the disease in humans and nonhuman primates. The Lyons research group has followed bacterial organ loads, disease time course, and pathology and found that i.n. infection is a valid pneumonic plague model in mice (C. R. Lyons, personal communication). Lathem et al. (54) calculated an i.n. LD50 for strain CO92 of 260 cells, with KIM5(pCD1Ap)+ giving a similar value. Our KIM5(pCD1Ap)+ strain yielded an i.n. instillation LD50 of 329 ± 105. At doses above the LD50, mice exhibited ruffled fur by day 2 postinfection, with deaths occurring between days 3 and 5. In contrast, our Δpsn and Δirp2 mutants yielded LD50s of approximately 104 and 105. Again we used two independent psn and irp2 mutations to ensure that the loss of virulence was not due to a secondary mutation (Table 2). While both the psn and irp2 mutants were significantly less virulent than WT, the difference between the two mutants was unexpected and intriguing. Table 2 shows the averaged LD50s for the strains unable to produce the Ybt siderophore (irp2 mutants) and those able to synthesize Ybt but unable to utilize it (psn mutants). An ∼24-fold difference (P = 0.0076) was maintained between the biosynthetic and transport mutants, with the psn mutants being 33-fold less virulent and the irp2 mutants being 790-fold less virulent than the WT. These results were surprising since in vitro, a biosynthetic mutant has less of an iron uptake defect than a mutant which produces but cannot utilize the Ybt siderophore (78). A time-to-death analysis of mice infected with the Ybt+ strain and the irp2 and psn mutants is shown in Fig. 3. Mice infected with the Ybt+ strain (average dose of 1,153 cells) died within 3 to 4 days postinfection. Higher average doses of the psn and irp2 mutants (17,050 and 268,000 cells, respectively) were lethal, beginning at 5 to 6 days postinfection and lasting until day 8.

FIG. 3.

FIG. 3.

Time-to-death analysis from i.n. instillation studies. Except for the Δpgm mutant, infectious doses used were close to the calculated LD50 for that strain. The average doses (in parentheses) were calculated from two (Ybt+ and psn), three (irp2), and four (Δpgm) independent experiments. All studies were carried out to 14 days, with no further deaths after day 10. Data are averages from all LD50 studies, shown as percent survival on the indicated days postinfection.

We previously hypothesized that the Ybt secreted by the transport mutant chelates residual iron in iron-deficient media, making it unavailable to other Y. pestis iron transport systems (78). Hence, these mutants would have more of a growth defect in iron-deficient media than strains that do not produce the siderophore. The growth patterns of biosynthetic and transport mutants of the Ybt system under iron-restrictive conditions are shown in Fig. 4. The transport (psn) mutant exhibits a significant growth defect compared to the WT strain and the biosynthetic mutant (Fig. 4). This growth defect is relieved in an irp2 psn mutant, which can neither synthesize nor use Ybt. Addition of the Ybt siderophore to the growth medium retards the growth of the double mutant (Fig. 4). Note that this growth defect is not as severe as observed with the psn mutant. This is likely due to the high concentration of Ybt which is produced by the psn mutant throughout the growth of the cells in deferrated PMH2 (three transfers); Ybt was added to the irp2 psn double mutant only during the third transfer shown in Fig. 4. Thus, these results indicate that, in vitro, Ybt production in a Ybt transport mutant is detrimental to iron-restricted growth.

FIG. 4.

FIG. 4.

Iron-deficient growth of Y. pestis strains. All strains were grown in deferrated PMH2 at 37°C. Where indicated, purified Ybt was added to KIM6+ (Ybt+) and KIM6-2180 (Δirp2-2046.3 Δpsn2045.1) at a concentration similar to that produced by a Ybt+ strain. The growth curves shown are from one of two independent experiments, which yielded similar results.

The pgm locus and plague.

Early studies showed that Pgm (putative or proven Δpgm) mutants were avirulent by peripheral routes of infection but fully virulent if injected intravenously (26, 52, 92). Specific ybt mutations within the pgm locus clearly demonstrate that the Ybt system is essential for bubonic plague and of critical importance for pneumonic plague (Table 2). We also tested two independent Δpgm mutants by an i.n. instillation route of infection. Both mutants yielded similar results; overall, 73.3% (11/15) of the mice survived administration of ∼3 × 106 cells. Thus, the LD50 of a Δpgm mutant is >11,850- and >15-fold higher than those of the WT and the irp2 mutant (P = 0.0013), respectively. At the highest dose, the Δpgm mutant killed fewer mice than the irp2 mutant and had a time-to-death range that was slightly delayed compared to that of the psn mutant (Fig. 3). If the avirulence of the Δpgm strain was strictly due to the absence of the Ybt iron transport system, then we would have expected it to have an LD50 and a time-to-death range similar to those of the irp2 mutant. These data clearly suggest that an additional factor or factors encoded within the pgm locus play a role in pneumonic plague.

DISCUSSION

Use of host iron sources by Ybt.

Our in vitro analyses have demonstrated that the Ybt siderophore can remove iron from Tf and have suggested that it can remove iron from Lf. A Ybt+ strain was capable of using these compounds as iron sources for growth when separated from the compounds by a dialysis baggy. Thus, a secreted diffusible molecule seems to be required for this growth response. We used gel electrophoresis to demonstrate that Ybt was involved in removing iron from Tf. In contrast, a Ybt biosynthetic mutant failed to respond to these iron sources when separated by a dialysis membrane, supporting the conclusion that the Ybt siderophore is required to use the bound iron under these conditions. The Tf and Lf results were not unexpected, since the Y. pestis KIM10+ genome (derived from KIM6+) encodes no apparent OM receptors for these compounds. Since we used iron chelators to prevent growth of Y. pestis strains without added host iron sources, it remains a formal possibility that the Ybt siderophore removed iron chelated by EDDA rather than directly from Lf.

Ybt and virulence.

Our previous LD50 studies of bubonic plague in mice used bacterial strains with background mutations in yopJ and psa that caused a slight attenuation compared to a fully virulent WT background (∼5-fold loss of virulence) (7, 8). To assess whether the yopJ psa background artificially enhanced the virulence defect due to ybt mutations, we tested both Ybt biosynthetic and transport mutants in an otherwise WT background. We also tested higher bacterial doses than in previous studies. Our results indicate that loss of the Ybt system causes a >4.3 × 105-fold loss of virulence by an s.c. route of infection. The Ybt biosynthetic and transport mutants showed similar decreases in virulence by this route of infection.

In other bacteria, siderophores as well as heme transport systems have been implicated in iron acquisition in the lung. For example, legiobactin, ornibactin, and alcaligin are required for lung infections by Legionella pneumophila, Burkholderia cenocepacia, and Bordetella pertussis, respectively (2, 19, 93). B. pertussis also uses enterobactin and a heme uptake system to acquire iron during lung infections (20, 22). In Klebsiella pneumoniae, Ybt played a major role in iron acquisition in the lung (55).

Two different mutations in psn (encoding the OM receptor for Ybt) caused an ∼33-fold loss of virulence. In contrast to our results in the bubonic model, two different irp2 mutants that are unable to produce the Ybt siderophore caused an even greater loss of virulence than the psn mutant, which can produce Ybt but is unable to use it: an ∼24-fold greater loss than with the psn mutants and a 790-fold loss of virulence compared to the parental WT strain (Table 2). Since the Ybt biosynthetic mutant was not completely avirulent, one of the other Y. pestis iron transport systems may be modestly effective in acquiring iron during a lung infection.

The difference in virulence between the psn and irp2 mutants is intriguing, especially since transport mutants are more detrimental to in vitro iron-restricted growth and iron uptake than biosynthetic mutants (78). This is the opposite of the results we found in the pneumonic plague model (i.e., the strain showing the most in vitro growth defect was more virulent). In the lungs, the Ybt siderophore may have other effects in addition to its role in providing iron. Although it is clear that Ybt serves as a signal molecule to activate transcription from ybt promoters (5, 36, 74, 77), the Ybt/YbtA signaling pathway in Y. pestis has not been entirely elucidated. We have favored a model in which the Ybt siderophore is transported into the cell and interacts with YbtA to transcriptionally activate regulated genes. There is good evidence for this type of regulation in Pseudomonas aeruginosa and Bordetella, which have AraC family regulators that respond to their cognate siderophores (17, 21, 67). While TonB-dependent signaling through the OM receptors of some other bacterial iron transport systems has been demonstrated (16), uptake mutants (psn, tonB, ybtP, and/or ybtQ) all show normal ybt gene regulation (7, 77, 80). The Ybt siderophore is a potent signaling molecule; growth stimulation by Ybt requires concentrations ∼500-fold higher than the concentration needed to activate transcription of the ybtP promoter (77). Consequently, we propose that small amounts of Ybt, sufficient to serve its signaling function, enter the cell via alternate routes. Since our in vitro studies indicate that a psn mutant can sense and respond to the Ybt siderophore, perhaps these small amounts are sufficient to activate transcription of other virulence factors in Y. pestis. In P. aeruginosa, the siderophore pyoverdine regulates not only its own production but also that of additional virulence determinants (9, 53).

An alternative to the signaling hypothesis is that the siderophore affects the host environment or innate immunity. Deferrated hydroxamate siderophores (desferrioxamine, desferrichrome, and desferriaerobactin) have an immunosuppressive effect on isolated mouse spleen mononuclear cells. In addition, enterobactin, independent of iron chelation, and desferrioxamine are cytotoxic for proliferating T cells (6, 49). However, other groups using desferrioxamine have found stimulatory effects on inflammatory cytokine production by intestinal and U937 cell lines (29, 57, 91). Pyochelin, which resembles Ybt structurally, can generate hydroxyl radicals and, under the appropriate conditions, damage pulmonary endothelial and epithelial cells (23, 24, 30). In addition, desferrithiocin, an iron chelator that is structurally similar to pyochelin, has been shown to inhibit T-cell proliferation (11). Recently, purified Ybt was shown to reduce the generation of reactive oxygen species by polymorphonuclear leukocytes (PMNs), human monocytes, and J774A.1 cells in vitro (71). Thus, Ybt may have direct toxic effects, affect host immune cell recruitment, and/or affect the synthesis of proinflammatory cytokines and/or reactive oxygen species.

The pgm locus and virulence.

Another unexpected finding of this study was that the Δpgm mutant showed an even greater loss of virulence than the irp2 mutants (>15-fold), a >11,800-fold loss of virulence compared to the WT strain (Table 2). Other investigators have examined the virulence of Δpgm strains via a pneumonic route of infection (73, 95). The LD50s obtained range from 104 to ∼106 cells. The differences in the observed LD50s could be the result of a number of factors, including the strain of mouse used, the method of administering the bacteria, and the way that the bacterial cells were grown. None of the other groups compared the virulence of a Δpgm strain with that of an Ybt mutant. Lee-Lewis and Anderson (58) did find that intraperitoneal administration of iron increased the virulence of a Δpgm strain but not to the level of a wild-type strain. Interestingly, this group also demonstrated that the Δpgm mutant did not cause pneumonic disease. The authors concluded that additional factors within the pgm locus play a role in pneumonic plague. Our results also clearly suggest that the 102-kb pgm locus, which includes the Ybt high-pathogenicity island and hms biofilm genes, encodes one or more virulence factors in addition to the Ybt system that play a role in pneumonic plague.

The only genes carried within the pgm locus for which there are published virulence studies are hmsR and hmsH; both are required for biofilm development (42, 60, 61, 75). An hmsR mutant was tested in a bubonic plague model, while the hmsH mutant was tested for virulence in mice via subcutaneous (bubonic plague) and intranasal (pneumonic plague) routes of infection. These studies found no significant role of biofilm formation in the virulence of either form of plague (1, 60).

The ripABC locus, also carried within the pgm locus, is required for survival in macrophages activated after bacterial infection (82). In collaboration with Jim Bliska's research group, we have found that the ΔripABC mutant was fully virulent in our mouse model of pneumonic plague (J. Bliska and R. D. Perry, unpublished observations). However, a number of other open reading frames (ORFs) that could have effects on virulence are carried within the remaining >60 kb of the pgm locus. There are loci potentially encoding a pilus, a ferrous transporter distantly related to the Ftr/Efe family and more closely related to a newly identified FetMP ferrous transporter (36, 47, 52a, 89), two cation transporters, and six transcriptional regulators. Extensive experimental analysis will be required to determine if one or more of these loci are involved in the further loss of virulence of the KIM Δpgm mutant (34) in our mouse model of pneumonic plague.

Acknowledgments

This study was supported by Public Health Service grant AI-33481 from the National Institutes of Health.

We thank our collaborator Jim Bliska for allowing us to report our findings on the rip mutation and virulence in pneumonic plague.

Editor: A. J. Bäumler

Footnotes

Published ahead of print on 16 February 2010.

REFERENCES

  • 1.Abu Khweek, A., J. D. Fetherston, and R. D. Perry. Analysis of HmsH and its role in plague biofilm formation. Microbiology, in press. [DOI] [PMC free article] [PubMed]
  • 2.Allard, K. A., J. Dao, P. Sanjeevaiah, K. McCoy-Simandle, C. H. Chatfield, D. S. Crumrine, D. Castignetti, and N. P. Cianciotto. 2009. Purification of legiobactin and the importance of this siderophore in lung infection by Legionella pneumophila. Infect. Immun. 77:2887-2895. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 3.Anderson, D. M., N. A. Ciletti, H. Lee-Lewis, D. Elli, J. Segal, K. L. DeBord, K. A. Overheim, M. Tretiakova, R. R. Brubaker, and O. Schneewind. 2009. Pneumonic plague pathogenesis and immunity in brown Norway rats. Am. J. Pathol. 174:910-921. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 4.Anderson, G. W., Jr., S. E. C. Leary, E. D. Williamson, R. W. Titball, S. L. Welkos, P. L. Worsham, and A. M. Friedlander. 1996. Recombinant V antigen protects mice against pneumonic and bubonic plague caused by F1-capsule-positive and -negative strains of Yersinia pestis. Infect. Immun. 64:4580-4585. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 5.Anisimov, R., D. Brem, J. Heesemann, and A. Rakin. 2005. Molecular mechanism of YbtA-mediated transcriptional regulation of divergent overlapping promoters ybtA and irp6 of Yersinia enterocolitica. FEMS Microbiol. Lett. 250:27-32. [DOI] [PubMed] [Google Scholar]
  • 6.Autenrieth, I., K. Hantke, and J. Heesemann. 1991. Immunosuppression of the host and delivery of iron to the pathogen: a possible dual role of siderophores in the pathogenesis of microbial infections. Med. Micriobiol. Immunol. 180:135-141. [DOI] [PubMed] [Google Scholar]
  • 7.Bearden, S. W., J. D. Fetherston, and R. D. Perry. 1997. Genetic organization of the yersiniabactin biosynthetic region and construction of avirulent mutants in Yersinia pestis. Infect. Immun. 65:1659-1668. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 8.Bearden, S. W., and R. D. Perry. 1999. The Yfe system of Yersinia pestis transports iron and manganese and is required for full virulence of plague. Mol. Microbiol. 32:403-414. [DOI] [PubMed] [Google Scholar]
  • 9.Beare, P. A., R. J. For, L. W. Martin, and I. L. Lamont. 2003. Siderophore-mediated cell signalling in Pseudomonas aeruginosa: divergent pathways regulate virulence factor production and siderophore receptor synthesis. Mol. Microbiol. 47:195-207. [DOI] [PubMed] [Google Scholar]
  • 10.Beesley, E. D., R. R. Brubaker, W. A. Janssen, and M. J. Surgalla. 1967. Pesticins. III. Expression of coagulase and mechanism of fibrinolysis. J. Bacteriol. 94:19-26. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 11.Bierer, B. E., and D. G. Nathan. 1990. The effect of desferrithiocin, an oral iron chelator, on T-cell function. Blood 76:2052-2059. [PubMed] [Google Scholar]
  • 12.Birnboim, H. C., and J. Doly. 1979. A rapid alkaline extraction procedure for screening recombinant plasmid DNA. Nucleic Acids Res. 7:1513-1523. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 13.Bobrov, A. G., V. A. Geoffroy, and R. D. Perry. 2002. Yersiniabactin production requires the thioesterase domain of HMWP2 and YbtD, a putative phosphopantetheinylate transferase. Infect. Immun. 70:4204-4214. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 14.Braun, V. 2005. Bacterial iron transport related to virulence. Contrib. Microbiol. 12:210-233. [DOI] [PubMed] [Google Scholar]
  • 15.Braun, V. 2001. Iron uptake mechanisms and their regulation in pathogenic bacteria. Int. J. Med. Microbiol. 291:67-79. [DOI] [PubMed] [Google Scholar]
  • 16.Braun, V., S. Mahren, and A. Sauter. 2006. Gene regulation by transmembrane signaling. BioMetals 19:103-113. [DOI] [PubMed] [Google Scholar]
  • 17.Brickman, T., and S. Armstrong. 2009. Temporal signaling and differential expression of Bordetella iron transport systems: the role of ferrimones and positive regulators. BioMetals 22:33-41. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 18.Brickman, T. J., and S. K. Armstrong. 2005. Bordetella AlcS transporter functions in alcaligin siderophore export and is central to inducer sensing in positive regulation of alcaligin system gene expression. J. Bacteriol. 187:3650-3661. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 19.Brickman, T. J., and S. K. Armstrong. 2007. Impact of alcaligin siderophore utilization on in vivo growth of Bordetella pertussis. Infect. Immun. 75:5305-5312. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 20.Brickman, T. J., T. Hanawa, M. T. Anderson, R. J. Suhadolc, and S. K. Armstrong. 2008. Differential expression of Bordetella pertussis iron transport system genes during infection. Mol. Microbiol. 70:3-14. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 21.Brickman, T. J., H. Y. Kang, and S. K. Armstrong. 2001. Transcriptional activation of Bordetella alcaligin siderophore genes requires the AlcR regulator with alcaligin as inducer. J. Bacteriol. 183:483-489. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 22.Brickman, T. J., C. K. Vanderpool, and S. K. Armstrong. 2006. Heme transport contributes to in vivo fitness of Bordetella pertussis during primary infection in mice. Infect. Immun. 74:1741-1744. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 23.Britigan, B. E., G. T. Rasmussen, and C. D. Cox. 1997. Augmentation of oxidant injury to human pulmonary epithelial cells by the Pseudomonas aeruginosa siderophore pyochelin. Infect. Immun. 65:1071-1076. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 24.Britigan, B. E., G. T. Rasmussen, and C. D. Cox. 1994. Pseudomonas siderophore pyochelin enhances neutrophil-mediated endothelial cell injury. Am. J. Physiol. 266:L192-L198. [DOI] [PubMed] [Google Scholar]
  • 25.Brubaker, R. R. 1969. Mutation rate to nonpigmentation in Pasteurella pestis. J. Bacteriol. 98:1404-1406. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 26.Brubaker, R. R., and M. J. Surgalla. 1961. Pesticins. I. Pesticin-bacterium interrelationships and environmental factors influencing activity. J. Bacteriol. 82:940-949. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 27.Bullen, J. J., H. J. Rogers, P. B. Spalding, and C. G. Ward. 2005. Iron and infection: the heart of the matter. FEMS Immunol. Med. Microbiol. 43:325-330. [DOI] [PubMed] [Google Scholar]
  • 28.Byrne, W. R., S. L. Welkos, M. L. Pitt, K. J. Davis, R. P. Brueckner, J. W. Ezzell, G. O. Nelson, J. R. Vaccaro, L. C. Battersby, and A. M. Friedlander. 1998. Antibiotic treatment of experimental pneumonic plague in mice. Antimicrob. Agents Chemother. 42:675-681. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 29.Choi, E.-Y., E.-C. Kim, H.-M. Oh, S. Kim, H.-J. Lee, E.-Y. Cho, K.-H. Yoon, E.-A. Kim, W.-C. Han, S.-C. Choi, J.-Y. Hwang, C. Park, B.-S. Oh, Y. Kim, K.-C. Kimm, K.-I. Park, H.-T. Chung, and C.-D. Jun. 2004. Iron chelator triggers inflammatory signals in human intestinal epithelial cells: involvement of p38 and extracellular signal-regulated kinase signaling pathways. J. Immunol. 172:7069-7077. [DOI] [PubMed] [Google Scholar]
  • 30.Coffman, T. J., C. D. Cox, B. L. Edeker, and B. E. Britigan. 1990. Possible role of bacterial siderophores in inflammation. Iron bound to the Pseudomonas siderophore pyochelin can function as a hydroxyl radical catalyst. J. Clin. Invest. 86:1030-1037. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 31.Crosa, J. H., A. R. Mey, and S. M. Payne (ed.). 2004. Iron transport in bacteria. ASM Press, Washington, DC.
  • 32.Dallaire, F., N. Ouellet, Y. Bergeron, V. Turmel, M. C. Gauthier, M. Simard, and M. G. Bergeron. 2001. Microbiological and inflammatory factors associated with the development of pneumococcal pneumonia. J. Infect. Dis. 184:292-300. [DOI] [PubMed] [Google Scholar]
  • 33.Davy, P., D. Bingham, G. Walters, and J. T. Whicher. 1982. Estimation of the saturation of serum transferrin by an electrophoretic technique. Ann. Clin. Biochem. 19:57-59. [DOI] [PubMed] [Google Scholar]
  • 34.Deng, W., V. Burland, G. Plunkett III, A. Boutin, G. F. Mayhew, P. Liss, N. T. Perna, D. J. Rose, B. Mau, S. Zhou, D. C. Schwartz, J. D. Fetherston, L. E. Lindler, R. R. Brubaker, G. V. Plano, S. C. Straley, K. A. McDonough, M. L. Nilles, J. S. Matson, F. R. Blattner, and R. D. Perry. 2002. Genome sequence of Yersinia pestis KIM. J. Bacteriol. 184:4601-4611. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 35.Dubbels, B. L., A. A. DiSpirito, J. D. Morton, J. D. Semrau, J. N. E. Neto, and D. A. Bazylinski. 2004. Evidence for a copper-dependent iron transport system in the marine, magnetotactic bacterium strain MV-1. Microbiology 150:2931-2945. [DOI] [PubMed] [Google Scholar]
  • 36.Fetherston, J. D., S. W. Bearden, and R. D. Perry. 1996. YbtA, an AraC-type regulator of the Yersinia pestis pesticin/yersiniabactin receptor. Mol. Microbiol. 22:315-325. [DOI] [PubMed] [Google Scholar]
  • 37.Fetherston, J. D., V. J. Bertolino, and R. D. Perry. 1999. YbtP and YbtQ: two ABC transporters required for iron uptake in Yersinia pestis. Mol. Microbiol. 32:289-299. [DOI] [PubMed] [Google Scholar]
  • 38.Fetherston, J. D., J. W. Lillard, Jr., and R. D. Perry. 1995. Analysis of the pesticin receptor from Yersinia pestis: role in iron-deficient growth and possible regulation by its siderophore. J. Bacteriol. 177:1824-1833. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 39.Fetherston, J. D., P. Schuetze, and R. D. Perry. 1992. Loss of the pigmentation phenotype in Yersinia pestis is due to the spontaneous deletion of 102 kb of chromosomal DNA which is flanked by a repetitive element. Mol. Microbiol. 6:2693-2704. [DOI] [PubMed] [Google Scholar]
  • 40.Fields, K. A., M. L. Nilles, C. Cowan, and S. C. Straley. 1999. Virulence role of V antigen of Yersinia pestis at the bacterial surface. Infect. Immun. 67:5395-5408. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 41.Ford, S., R. A. Cooper, R. W. Evans, R. C. Hider, and P. H. Williams. 1988. Domain preference in iron removal from human transferrin by the bacterial siderophores aerobactin and enterochelin. Eur. J. Biochem. 178:477-481. [DOI] [PubMed] [Google Scholar]
  • 42.Forman, S., A. G. Bobrov, O. Kirillina, S. K. Craig, J. Abney, J. D. Fetherston, and R. D. Perry. 2006. Identification of critical amino acid residues in the plague biofilm Hms proteins. Microbiology 152:3399-3410. [DOI] [PubMed] [Google Scholar]
  • 43.Forman, S., M. J. Nagiec, J. Abney, R. D. Perry, and J. D. Fetherston. 2007. Analysis of the aerobactin and ferric hydroxamate uptake systems of Yersinia pestis. Microbiology 153:2332-2341. [DOI] [PubMed] [Google Scholar]
  • 44.Furrer, J. L., D. N. Sanders, I. G. Hook-Barnard, and M. A. McIntosh. 2002. Export of the siderophore enterobactin in Escherichia coli: involvement of a 43 kDa membrane exporter. Mol. Microbiol. 44:1225-1234. [DOI] [PubMed] [Google Scholar]
  • 45.Gong, S., S. W. Bearden, V. A. Geoffroy, J. D. Fetherston, and R. D. Perry. 2001. Characterization of the Yersinia pestis Yfu ABC iron transport system. Infect. Immun. 67:2829-2837. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 46.Groβe, C., J. Scherer, D. Koch, M. Otto, N. Taudte, and G. Grass. 2006. A new ferrous iron-uptake transporter, EfeU (YcdN), from Escherichia coli. Mol. Microbiol. 62:120-131. [DOI] [PubMed] [Google Scholar]
  • 47.Heinrichs, D. E., and K. Poole. 1993. Cloning and sequence analysis of a gene (pchR) encoding an AraC family activator of pyochelin and ferripyochelin receptor synthesis in Pseudomonas aeruginosa. J. Bacteriol. 175:5882-5889. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 48.Higuchi, K., and J. L. Smith. 1961. Studies on the nutrition and physiology of Pasteurella pestis. VI. A differential plating medium for the estimation of the mutation rate to avirulence. J. Bacteriol. 81:605-608. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 49.Hileti, D., P. Panayiotidis, and A. V. Hoffbrand. 1995. Iron chelators induce apoptosis in proliferating cells. Br. J. Haematol. 89:181-187. [DOI] [PubMed] [Google Scholar]
  • 50.Hissen, A. H. T., J. M. T. Chow, L. J. Pinto, and M. M. Moore. 2004. Survival of Aspergillus fumigatus in serum involves removal of iron from transferrin: the role of siderophores. Infect. Immun. 72:1402-1408. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 51.Hoover, D. L., R. M. Crawford, L. L. Van De Verg, M. J. Izadjoo, A. K. Bhattacharjee, C. M. Paranavitana, R. L. Warren, M. P. Nikolich, and T. L. Hadfield. 1999. Protection of mice against brucellosis by vaccination with Brucella melitensis WR201(16MDpurEK). Infect. Immun. 67:5877-5884. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 52.Jackson, S., and T. W. Burrows. 1956. The virulence-enhancing effect of iron on non-pigmented mutants of virulent strains of Pasteurella pestis. Br. J. Exp. Pathol. 37:577-583. [PMC free article] [PubMed] [Google Scholar]
  • 52a.Koch, D., D. H. Nies, and G. Grass. 2008. Characterization of a novel iron uptake system from uropathogenic Escherichia coli strain F11. Abstr. BioMetals 2008, p. 102. [DOI] [PMC free article] [PubMed]
  • 53.Lamont, I. L., P. A. Beare, U. Ochsner, A. I. Vasil, and M. L. Vasil. 2002. Siderophore-mediated signaling regulates virulence factor production in Pseudomonas aeruginosa. Proc. Natl. Acad. Sci. U. S. A. 99:7072-7077. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 54.Lathem, W. W., S. D. Crosby, V. L. Miller, and W. E. Goldman. 2005. Progression of primary pneumonic plague: a mouse model of infection, pathology, and bacterial transcriptional activity. Proc. Natl. Acad. Sci. U. S. A. 102:17786-17791. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 55.Lawlor, M. S., C. O'Connor, and V. L. Miller. 2007. Yersiniabactin is a virulence factor for Klebsiella pneumoniae during pulmonary infection. Infect. Immun. 75:1463-1472. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 56.Lawson, J. N., C. R. Lyons, and S. A. Johnston. 2006. Expression profiling of Yersinia pestis during mouse pulmonary infection. DNA Cell Biol. 25:608-616. [DOI] [PubMed] [Google Scholar]
  • 57.Lee, H. J., S. C. Choi, E. Y. Choi, M. H. Lee, G. S. Seo, E. C. Kim, B. J. Yang, M. S. Lee, Y. I. Shin, K. I. Park, and C. D. Jun. 2005. Iron chelator induces MIP-α/CCL20 in human intestinal epithelial cells: implication for triggering mucosal adaptive immunity. Exp. Mol. Med. 37:297-310. [DOI] [PubMed] [Google Scholar]
  • 58.Lee-Lewis, H., and D. M. Anderson. 2010. Absence of inflammation and pneumonia during infection with nonpigmented Yersinia pestis reveals new role for the pgm locus in pathogenesis. Infect. Immun. 78:220-230. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 59.Lesic, B., and E. Carniel. 2004. The high pathogenicity island: a broad-host-range pathogenicity island, p. 285-306. In E. Carniel and B. J. Hinnebusch (ed.), Yersinia molecular and cellular biology. Horizon Bioscience, Norfolk, United Kingdom.
  • 60.Lillard, J. W., Jr., S. W. Bearden, J. D. Fetherston, and R. D. Perry. 1999. The haemin storage (Hms+) phenotype of Yersinia pestis is not essential for the pathogenesis of bubonic plague in mammals. Microbiology 145:197-209. [DOI] [PubMed] [Google Scholar]
  • 61.Lillard, J. W., Jr., J. D. Fetherston, L. Pedersen, M. L. Pendrak, and R. D. Perry. 1997. Sequence and genetic analysis of the hemin storage (hms) system of Yersinia pestis. Gene 193:13-21. [DOI] [PubMed] [Google Scholar]
  • 62.Lucier, T. S., and R. R. Brubaker. 1992. Determination of genome size, macrorestriction pattern polymorphism, and nonpigmentation-specific deletion in Yersinia pestis by pulsed-field gel electrophoresis. J. Bacteriol. 174:2078-2086. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 63.Lynch, D., J. O'Brien, T. Welch, P. Clarke, P. O Cuiv, J. H. Crosa, and M. O'Connell. 2001. Genetic organization of the region encoding regulation, biosynthesis, and transport of rhizobactin 1021, a siderophore produced by Sinorhizobium meliloti. J. Bacteriol. 183:2576-2585. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 64.Makey, D. G., and U. S. Seal. 1976. The detection of four molecular forms of human transferrin during the iron binding process. Biochim. Biophys. Acta 453:250-256. [DOI] [PubMed] [Google Scholar]
  • 65.Matthijs, S., C. Baysse, N. Koedam, K. A. Tehrani, L. Verheyden, H. Budzikiewicz, M. Schäfer, B. Hoorelbeke, J.-M. Meyer, H. De Greve, and P. Cornelis. 2004. The Pseudomonas siderophore quinolobactin is synthesized from xanthurenic acid, an intermediate of the kynurenine pathway. Mol. Microbiol. 52:371-384. [DOI] [PubMed] [Google Scholar]
  • 66.Michel, L., A. Bachelard, and C. Reimmann. 2007. Ferripyochelin uptake genes are involved in pyochelin-mediated signalling in Pseudomonas aeruginosa. Microbiology 153:1508-1518. [DOI] [PubMed] [Google Scholar]
  • 67.Michel, L., N. González, S. Jagdeep, T. Nguyen-Ngoc, and C. Reimmann. 2005. PchR-box recognition by the AraC-type regulator PchR of Pseudomonas aeruginosa requires the siderophore pyochelin as an effector. Mol. Microbiol. 58:495-509. [DOI] [PubMed] [Google Scholar]
  • 68.Miller, M. C., S. Parkin, J. D. Fetherston, R. D. Perry, and E. DeMoll. 2006. Crystal structure of ferric-yersiniabactin, a virulence factor of Yersinia pestis. J. Inorg. Biochm. 100:1495-1500. [DOI] [PubMed] [Google Scholar]
  • 69.Morales, S. E., and T. A. Lewis. 2006. Transcriptional regulation of the pdt gene cluster of Pseudomonas stutzeri KC involves an AraC/XylS family transcriptional activator (PdtC) and the cognate siderophore pyridine-2,6-bis(thiocarboxylic acid). Appl. Environ. Microbiol. 72:6994-7002. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 69a.National Institutes of Health. 2000. Public Health Service policy on humane care and use of laboratory animals. Office of Laboratory Animal Welfare, National Institutes of Health, Bethesda, MD.
  • 69b.National Research Council. 1996. Guide for the care and use of laboratory animals. National Academy Press, Washington, DC.
  • 70.Ó Cuív, P., P. Clarke, D. Lynch, and M. O'Connell. 2004. Identification of rhtX and fptX, novel genes encoding proteins that show homology and function in the utilization of the siderophores rhizobactin 1021 by Sinorhizobium meliloti and pyochelin by Pseudomonas aeruginosa, respectively. J. Bacteriol. 186:2996-3005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 70a.Office of Science and Technology Policy. 1985. U.S. Government principles for the utilization and care of vertebrate animals used in teaching, research, and training. Fed. Regist. 50:20864-20865. [PubMed] [Google Scholar]
  • 71.Paauw, A., M. A. Leverstein-van Hall, K. P. M. van Kessel, J. Verhoef, and A. C. Fluit. 2009. Yersiniabactin reduces the respiratory oxidative stress response of innate immune cells. PLoS One 4:e8240. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 72.Pal, S., E. M. Peterson, and L. M. de La Maza. 2000. Role of Nramp1 deletion in Chlamydia infection in mice. Infect. Immun. 68:4831-4833. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 73.Parent, M. A., K. N. Berggren, L. W. Kummer, L. B. Wilhelm, F. M. Szaba, I. K. Mullarky, and S. T. Smiley. 2005. Cell-mediated protection against pulmonary Yersinia pestis infection. Infect. Immun. 73:7304-7310. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 74.Pelludat, C., A. Rakin, C. A. Jacobi, S. Schubert, and J. Heesemann. 1998. The yersiniabactin biosynthetic gene cluster of Yersinia enterocolitica: organization and siderophore-dependent regulation. J. Bacteriol. 180:538-546. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 75.Pendrak, M. L., and R. D. Perry. 1993. Proteins essential for expression of the Hms+ phenotype of Yersinia pestis. Mol. Microbiol. 8:857-864. [DOI] [PubMed] [Google Scholar]
  • 76.Perry, R. D. 2004. Yersinia, p. 219-240. In J. H. Crosa, A. R. Mey, and S. M. Payne (ed.), Iron transport in bacteria. ASM Press, Washington, DC.
  • 77.Perry, R. D., J. Abney, I. Mier, Jr., Y. Lee, S. W. Bearden, and J. D. Fetherston. 2003. Regulation of the Yersinia pestis Yfe and Ybt iron transport systems. Adv. Exp. Med. Biol. 529:275-283. [DOI] [PubMed] [Google Scholar]
  • 78.Perry, R. D., P. B. Balbo, H. A. Jones, J. D. Fetherston, and E. DeMoll. 1999. Yersiniabactin from Yersinia pestis: biochemical characterization of the siderophore and its role in iron transport and regulation. Microbiology 145:1181-1190. [DOI] [PubMed] [Google Scholar]
  • 79.Perry, R. D., and J. D. Fetherston. 2004. Iron and heme uptake systems, p. 257-283. In E. Carniel and B. J. Hinnebusch (ed.), Yersinia molecular and cellular biology. Horizon Bioscience, Norfolk, United Kingdom.
  • 80.Perry, R. D., J. Shah, S. W. Bearden, J. M. Thompson, and J. D. Fetherston. 2003. Yersinia pestis TonB: role in iron, heme, and hemoprotein utilization. Infect. Immun. 71:4159-4162. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 81.Priebe, G. P., G. J. Meluleni, F. T. Coleman, J. B. Goldberg, and G. B. Pier. 2003. Protection against fatal Pseudomonas aeruginosa pneumonia in mice after nasal immunization with a live, attenuated aroA deletion mutant. Infect. Immun. 71:1453-1461. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 82.Pujol, C., J. P. Grabenstein, R. D. Perry, and J. B. Bliska. 2005. Replication of Yersinia pestis in interferon γ-activated macrophages requires ripA, a gene encoded in the pigmentation locus. Proc. Natl. Acad. Sci. U. S. A. 102:12909-12914. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 83.Putman, M., H. W. van Veen, and W. N. Konings. 2000. Molecular properties of bacterial multidrug transporters. Microbiol. Mol. Biol. Rev. 64:672-693. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 84.Reed, L. J., and H. Muench. 1938. A simple method for estimating fifty percent endpoints. Am. J. Hyg. 27:493-497. [Google Scholar]
  • 85.Schaible, U. E., and S. H. E. Kaufmann. 2004. Iron and microbial infection. Nat. Rev. Microbiol. 2:946-953. [DOI] [PubMed] [Google Scholar]
  • 86.Staggs, T. M., and R. D. Perry. 1991. Identification and cloning of a fur regulatory gene in Yersinia pestis. J. Bacteriol. 173:417-425. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 87.Steijns, J. M., and A. C. van Hooijdonk. 2000. Occurrence, structure, biochemical properties and technological characteristics of lactoferrin. Br. J. Nutr. 84(Suppl. 1):S11-S17. [DOI] [PubMed] [Google Scholar]
  • 88.Surgalla, M. J., and E. D. Beesley. 1969. Congo red-agar plating medium for detecting pigmentation in Pasteurella pestis. Appl. Microbiol. 18:834-837. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 89.Szczepanowski, R., S. Braun, V. Riedel, S. Schneiker, I. Krahn, A. Pühler, and A. Schlüter. 2005. The 120 592 bp IncF plasmid pRSB107 isolated from a sewage-treatment plant encodes nine different antibiotic-resistance determinants, two iron-acquisition systems and other putative virulence-associated functions. Microbiology 151:1095-1111. [DOI] [PubMed] [Google Scholar]
  • 90.Takase, H., H. Nitanai, K. Hoshino, and T. Otani. 2000. Impact of siderophore production on Pseudomonas aeruginosa infections in immunosuppressed mice. Infect. Immun. 68:1834-1839. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 91.Tanji, K., T. Imaizumi, T. Matsumiya, H. Itaya, K. Fujimoto, X.-f. Cui, T. Toki, E. Ito, H. Yoshida, K. Wakabayashi, and K. Satoh. 2001. Desferrioxamine, an iron chelator, upregulates cyclooxygenase-2 expression and prostaglandin production in a human macrophage cell line. Biochim. Biophys.Acta 1530:227-235. [DOI] [PubMed] [Google Scholar]
  • 92.Une, T., and R. R. Brubaker. 1984. In vivo comparison of avirulent Vwa and Pgm or Pstr phenotypes of yersiniae. Infect. Immun. 43:895-900. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 93.Visser, M. B., S. Majumdar, E. Hani, and P. A. Sokol. 2004. Importance of the ornibactin and pyochelin siderophore transport systems in Burkholderia cenocepacia lung infections. Infect. Immun. 72:2850-2857. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 94.Walsh, C. T., and C. G. Marshall. 2004. Siderophore biosynthesis in bacteria, p. 18-37. In J. H. Crosa, A. R. Mey, and S. M. Payne (ed.), Iron transport in bacteria. ASM Press, Washington, DC.
  • 95.Welkos, S., M. L. M. Pitt, M. Martinez, A. Friedlander, P. Vogel, and R. Tammariello. 2002. Determination of the virulence of the pigmentation-deficient and pigmentation-/plasminogen activator-deficient strains of Yersinia pestis in non-human primate and mouse models of pneumonic plague. Vaccine 20:2206-2214. [DOI] [PubMed] [Google Scholar]
  • 96.Wolz, C., K. Hohloch, A. Ocaktan, K. Poole, R. W. Evans, N. Rochel, A.-M. Albrecht-Gary, M. A. Abdallah, and G. Döring. 1994. Iron release from transferrin by pyoverdin and elastase from Pseudomonas aeruginosa. Infect. Immun. 62:4021-4027. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 97.Wu, T. H., J. A. Hutt, K. A. Garrison, L. S. Berliba, Y. Zhou, and C. R. Lyons. 2005. Intranasal vaccination induces protective immunity against intranasal infection with virulent Francisella tularensis biovar A. Infect. Immun. 73:2644-2654. [DOI] [PMC free article] [PubMed] [Google Scholar]

Articles from Infection and Immunity are provided here courtesy of American Society for Microbiology (ASM)

RESOURCES