ArticlePDF Available

The Moraxella catarrhalis AdhC–FghA system is important for formaldehyde detoxification and protection against pulmonary clearance

Authors:

Abstract and Figures

Multidrug-resistant clinical isolates of Moraxella catarrhalis have emerged, increasing the demand for the identification of new treatment and prevention strategies. A thorough understanding of how M. catarrhalis can establish an infection and respond to different stressors encountered in the host is crucial for new drug-target identification. Formaldehyde is a highly cytotoxic compound that can be produced endogenously as a by-product of metabolism and exogenously from environmental sources. Pathways responsible for formaldehyde detoxification are thus essential and are found in all domains of life. The current work investigated the role of the system consisting of the S-hydroxymethyl alcohol dehydrogenase (AdhC), a Zn-dependent class III alcohol dehydrogenase, and the S-formyl glutathione hydrolase (FghA) in the formaldehyde detoxification process in M. catarrhalis. Bioinformatics showed that the components of the system are conserved across the species and are highly similar to those of Streptococcus pneumoniae, which share the same biological niche. Isogenic mutants were constructed to study the function of the system in M. catarrhalis. A single fghA knockout mutant did not confer sensitivity to formaldehyde, while the adhC–fghA double mutant is formaldehyde-sensitive. In addition, both mutants were significantly cleared in a murine pulmonary model of infection as compared to the wild type, demonstrating the system’s importance for this pathogen’s virulence. The respective phenotypes were reversed upon the genetic complementation of the mutants. To date, this is the first study investigating the role of the AdhC–FghA system in formaldehyde detoxification and pathogenesis of M. catarrhalis.
This content is subject to copyright. Terms and conditions apply.
Vol.:(0123456789)
Medical Microbiology and Immunology (2024) 213:3
https://doi.org/10.1007/s00430-024-00785-0
ORIGINAL INVESTIGATION
The Moraxella catarrhalis AdhC–FghA system isimportant
forformaldehyde detoxification andprotection againstpulmonary
clearance
DinaOthman1· NohaM.Elhosseiny2· WafaaN.Eltayeb3· AhmedS.Attia2
Received: 21 October 2023 / Accepted: 24 January 2024
© The Author(s) 2024
Abstract
Multidrug-resistant clinical isolates of Moraxella catarrhalis have emerged, increasing the demand for the identification
of new treatment and prevention strategies. A thorough understanding of how M. catarrhalis can establish an infection and
respond to different stressors encountered in the host is crucial for new drug-target identification. Formaldehyde is a highly
cytotoxic compound that can be produced endogenously as a by-product of metabolism and exogenously from environmental
sources. Pathways responsible for formaldehyde detoxification are thus essential and are found in all domains of life. The
current work investigated the role of the system consisting of the S-hydroxymethyl alcohol dehydrogenase (AdhC), a Zn-
dependent class III alcohol dehydrogenase, and the S-formyl glutathione hydrolase (FghA) in the formaldehyde detoxifica-
tion process in M. catarrhalis. Bioinformatics showed that the components of the system are conserved across the species
and are highly similar to those of Streptococcus pneumoniae, which share the same biological niche. Isogenic mutants were
constructed to study the function of the system in M. catarrhalis. A single fghA knockout mutant did not confer sensitivity to
formaldehyde, while the adhC–fghA double mutant is formaldehyde-sensitive. In addition, both mutants were significantly
cleared in a murine pulmonary model of infection as compared to the wild type, demonstrating the system’s importance for
this pathogen’s virulence. The respective phenotypes were reversed upon the genetic complementation of the mutants. To
date, this is the first study investigating the role of the AdhC–FghA system in formaldehyde detoxification and pathogenesis
of M. catarrhalis.
Keywords Moraxella catarrhalis· Formaldehyde resistance· S-hydroxymethyl alcohol dehydrogenase· S-formyl
glutathione hydrolase· Pulmonary clearance
Introduction
Moraxella catarrhalis, previously considered a commensal
microorganism, has been commonly implicated as the major
etiological agent of otitis media and sinusitis in children,
and in exacerbation of chronic obstructive pulmonary dis-
ease in adults [1]. Multidrug-resistant clinical isolates of
M. catarrhalis have emerged, increasing the demand for
the identification of new treatment and prevention strate-
gies against this pathogen, especially in the absence of an
efficient vaccine [2, 3].
A formaldehyde detoxification system is essential for
microorganisms to protect themselves from cytotoxic formal-
dehyde [4]. In the human body, formaldehyde is produced
during the metabolism of methanol, adrenaline, creatine, and
histones. Most importantly, it is a by-product of some chemi-
cal reactions associated with the immune response, such as
Edited by: Isabelle Bekeredjian-Ding.
* Ahmed S. Attia
ahmed.attia@pharma.cu.edu.eg
1 Graduate Program, Department ofMicrobiology
andImmunology, Faculty ofPharmacy, Cairo University,
Cairo11562, Egypt
2 Department ofMicrobiology andImmunology, Faculty
ofPharmacy, Cairo University, Room #D404, Kasr El-Ainy
Street, Cairo11562, Egypt
3 Department ofMicrobiology, Faculty ofPharmacy, Misr
International University, Cairo19648, Egypt
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
Medical Microbiology and Immunology (2024) 213:3 3 Page 2 of 10
the methylation of histamine. During the respiratory burst by
macrophages and neutrophils to kill pathogens, superoxide and
hydrogen peroxide are generated, which in turn react with bac-
terial iron-sulfur clusters to produce free radicals. These free
radicals react with sugars to produce formaldehyde as a toxic
end product that could be used to fight invading pathogens [4].
Formaldehyde is also produced as a part of bacterial metabo-
lism in cell biological processes [5]. For all of the aforemen-
tioned reasons, the formaldehyde detoxification processes are
required to avoid formaldehyde lethal and mutagenic effect [6].
Formaldehyde detoxification is achieved through three
mechanisms; thiol-dependent, ribulose monophosphate-
dependent, and pterin-dependent mechanisms [4]. The glu-
tathione-dependent repair system, also known as the thiol-
dependent pathway, appears to be widely spread in nature
and has been found in most prokaryotes, and all eukaryotes
[7]. In the majority of microorganisms, the thiol is the trip-
eptide glutathione. Initially, glutathione binds to formalde-
hyde to formS-hydroxymethylglutathione [8]. This reaction
occurs spontaneously in most microorganisms, with some
exceptions where this reaction is catalyzed by a glutathione-
dependent formaldehyde-activatingenzyme, Gfa [9, 10].
The S-hydroxymethylglutathione adduct is then oxidized
by a zinc-containing, nicotinamide adenine dinucleotide
(NAD+)-dependent alcohol dehydrogenase, AdhC, to gener-
ate the thioesterS-formylglutathione [11]. Finally, formate is
produced and glutathione is regenerated upon the hydrolysis of
S-formyl glutathione. This last step is catalyzed by an esterase,
EstD, orS-formylglutathione hydrolase, FghA [12].
Numerous studies have been investigating the genetic fac-
tors involved in formaldehyde detoxification and their role in
stress protection, bacterial virulence, and biofilm formation
in different microorganisms [6, 7, 1319]. For example, the
glutathione-dependent formaldehyde detoxification system
AdhC–EstD was shown to be important for the optimum via-
bility of Neisseria meningitidis in biofilm communities [20]. In
another study, the loss of the encoding gene of EstD in N. gon-
orrhoeae caused an impairment in the ability of this organism
to survive within human cervical epithelial cells [17]. While
many of the genetic factors contributing to the physiology and
virulence of M. catarrhalis have been identified [2123], the
role of formaldehyde detoxification in these processes has not
been yet investigated. In this study, the adhC and fghA genes
of M. catarrhalis have been investigated with respect to their
potential role in the formaldehyde detoxification and the viru-
lence of this pathogen.
Materials andmethods
Ethics statement
Animal procedures were approved by the Research Ethics
Committee of the Faculty of Pharmacy, Cairo University,
Approval No. MI (2510), following the Guide for the Care
and Use of Laboratory Animals published by the Institute
of Laboratory Animal Research, USA.
Bacterial strains andculture conditions
M. catarrhalis O35E was kindly provided by Dr. Eric
J. Hansen [24] and it was used as the wild type (WT).
The derivatives were all generated in its background. M.
catarrhalis strains were grown on Tryptic Soy Agar (TSA)
or Columbia blood agar at 37°C and 5% CO2 in a carbon
dioxide incubator (Binder, Germany), or in Tryptic Soy
Broth (TSB) at 37°C with shaking at 180rpm under aerobic
conditions [25]. Escherichia coli DH5-α, used as a cloning
host, was grown at 37°C in TSB with shaking at 180rpm,
or on TSA. When needed, media were supplemented with
kanamycin at a final concentration of 15µg/mL, streptomy-
cin at a final concentration of 250µg/mL, and ampicillin at
a final concentration of 100μg/mL.
Bioinformatics analyses
The protein sequenceof the EstD of N. meningitidis (NCBI
protein id CAM08666.1) served as a template for a BlastP
analysis [26], to identify homologs of this esterase inM.
catarrhalis O35E.To survey the conservation of the pro-
tein, the BlastP search was also extended to all the strains
available in M. catarrhalis taxid 480. Sequences of pre-
viously studied FghA homologs were retrieved from the
NCBI (TableS1), and the same was performed with the
AdhC homologs (TableS2). Then, multiple sequences’
alignment with the Blast-retrieved FghA of M. catarrhalis
O35E (NCBI Protein id EGE27440.1) as a query sequence
was carried on using Clustal Omega [27] applyling the
default parameters. Phylogenetic trees were constructed via
NGphylogeny.fr web tool, which employs multiple align-
ment using fast Fourier transform (MAFFT) for multiple
sequence alignment, Block Mapping and Gathering with
Entropy (BMGE) for alignment curation, Phylogeny soft-
ware for the maximum-likelihood principle (PhyML) for
tree inference, and finally Newick display for tree render-
ing [2832]. To calculate the percent similarity and identity,
EMBOSS Needle tool was used with the default parameters
[33]. To further confirm the interaction between AdhC and
FghA and show possible interactions with other proteins, the
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
Medical Microbiology and Immunology (2024) 213:3 Page 3 of 10 3
protein–protein functional interaction analysis tool STRING
[34] was used. To investigate if the adhC and fghA genes
form an operonic pair as consistently reported for the sys-
tem, the operon prediction tool Operon Mapper [35] was
used applying the default parameters, and the whole M.
catarrhalis O35E genome (AERL00000000.1) as a query
sequence.
Construction ofM. catarrhalis ΔfghAand ΔadhC–
fghA deletion mutants andrescue strains
Using theM. catarrhalisO35E chromosomal DNA as a tem-
plate, the primer pairs DO001–DO002 and DO003–DO004
(TableS3 and Fig.1) were used to amplify a 1002 and a 1006
base pair fragments upstream and downstream of thefghA
open-reading frame (ORF), respectively. The fragments were
then digested using XmaIand ligated together. The ligation
product was used as a template for a second PCR reaction
using primer pair DO001–DO004. The product was then
ligated into the rapid cloning vector pJET1.2/blunt (Thermo
Fisher Scientific, Lithuania), to yield plasmid pJET-U + D.
A non-polar kanamycin resistance cassette was obtained by
digesting plasmid pUC18K [36] with XmaI, followed by
gel purification. The product was then ligated with plasmid
pJET-U + D digested with the same restriction enzyme, then
transformed into E. coli DH5-α and plated on TSA contain-
ing ampicillin and kanamycin. The resultant plasmid was
designated pJET-UkanD. This plasmid was used to amplify
a ~ 3kb construct consisting of the kanamycin cassette
flanked by the fghA upstream and downstream fragments,
and the product was transformed intoM. catarrhalisO35E
as previously described [37]. The homologous recombina-
tion of the mutant construct into the chromosome and the
replacement of the WT fghA gene to yield ΔfghAmutant
was confirmed by a series of PCR reactions using primers
within and outside the mutant construct. To construct the
ΔadhC-fghA double mutant,a similar approach was adopted
but using primer pairs DO008–DO009 and DO003–DO004
Fig. 1 Genetic organization and functional relation between the M.
catarrhalis adhC and fghA. A A schematic diagram showing the
organization of the neighboring ORFs in the genetic loci of the adhC
and fghA in the M. catarrhalis WT O35E genome. The binding posi-
tion of the primers used in this study and the loci tags are indicated.
The direction of the ORF arrows indicates the direction of transcrip-
tion. The ruler above the arrows indicates the size of DNA fragment
in base pairs; bp. The map was generated by Ankh diagram v1.1tool
by HITS Solutions Co. (Bioinformatics Department, Cairo, Egypt).
B A schematic diagram representing the interaction between AdhC
(encoded by EA1_02212) with FghA and other M. catarrhalis pro-
teins including Gdsl-like lipase (encoded by EA1_06621), MsrAB,
and GdhA. The figure was generated using STRING database and the
lines drawn between the functional pairs represent the predicted func-
tional relationship (neighborhood; green, gene fusion; red, co-occur-
rence; dark purple, co-expression; black, databases; teal, text mining;
yellow, experimentally determined interactions; pink line, and protein
homology; light blue)
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
Medical Microbiology and Immunology (2024) 213:3 3 Page 4 of 10
to amplify a 329 and a 1002 base pair fragments upstream
and downstream of adhC and fghA, respectively. To repair
the ΔfghAmutant, the fragment amplified with primer pair
DO001–DO004 from the WT O35E M. catarrhalis strain
together with the mutatedrpsLamplicon [38] were trans-
formed into theΔfghAin a congression experiment as pre-
viously described [39]. Isolated colonies that could grow
on streptomycin, and failed to grow on kanamycin, were
selected as potential complemented mutants, confirmed
using PCR, and designated ΔfghA/R. To repair ΔadhC–fghA
mutant, a similar approach was used, but using primer pair
DO008–DO004 to amplify the WTamplicon. The con-
firmed rescue mutant was designated ΔadhC–fghA/R. The
sequences of all the oligonucleotides used in this study are
listed in TableS3.
Growth curve analysis
Colonies of the WT O35E, ΔfghA, ΔfghA/R, ΔadhC-fghA,
and ΔadhC-fghA/Rgrown overnight on Columbia blood
agar were suspended in TSB to an optical density at 600nm
(OD600) ~ 1.0 then diluted 1:50 in 15mL TSB broth. The
cultures were incubated in a shaking incubator at 37°C and
180rpm and the OD600was measured each hour for 8h
using a visible spectrophotometer Jenway 6300 (Jenway,
United Kingdom). Growth curves were constructed by plot-
ting OD600 versus time.
Formaldehyde sensitivity assay
Formaldehyde susceptibility was assessed using the disc
diffusion susceptibility assay as previously described [20].
Briefly, cells of the WT O35E, ΔfghA, ΔfghA/R, ΔadhC-
fghA, and ΔadhC-fghA/Rfreshly streaked on Columbia
blood agar were suspended in TSB to an OD600 ~ 0.4. The
adjusted cell suspensions were spread over TSA plates
(supplemented with kanamycin for ΔfghA and ΔadhC-fghA
mutants, streptomycin for ΔfghA/R and ΔadhC-fghA/Rcom-
plemented mutants) using sterile cotton swabs. Sterile single
Whatmann no. 1 filter paper discs were saturated with 5 μL
of a 5% formaldehyde solution (Piochem, Egypt) and placed
onto the agar surface. The diameter of the inhibition zone
around the disc was measured after an overnight incubation
at 37°C in a CO2 incubator with the petri-dishes inverted.
Formaldehyde sensitivity was also tested by another assay
[20]. Briefly, a bacterial suspension was prepared as in the
sensitivity assay detailed above, and seven tenfold serial
dilutions were prepared in a 96-well microplate in TSB. Five
microliters of each dilution were spotted on TSA plates, sup-
plemented with 0-, 0.8-, or 1-mM formaldehyde. The spots
were left to dry then survival was determined by counting
the number of visible colonies after an overnight incubation
at 37°C in a CO2 incubator with the petri-dishes inverted.
Determination oftheminimum inhibitory
concentration (MIC)
To obtain a more quantitative assessment of the susceptibility
of the five strains under investigation to formaldehyde, we per-
formed an MIC experiment using the standard microdilution
method [40]. Briefly, a 0.5 McFarland standard suspension
of each of the five strains was prepared using freshly grown
bacterial cells, and then, it was diluted 1:10 and 10 µL were
used to inoculate 12 wells containing 190 µL of TSB contain-
ing twofold dilutions of formaldehyde from 1mM.to 0.5µM.
The wells incubated for 24h at 37°C then inspected for
visual growth. The formaldehyde concentration in the first
clear well was considered the corresponding MIC value for
the respective strain.
Protein profiling using sodium dodecyl
sulfate‑polyacrylamide gel electrophoresis
(SDS‑PAGE)
Cells of the five tested strains were grown overnight in TSB
in the presence and absence of 1mM formaldehyde in a shak-
ing incubator, harvested by centrifugation, and resuspended
in sterile saline to an OD600 ~ 1. The cell suspensions were
mixed with a 3 × reducing Laemmli buffer [41]. These sam-
ples were then heated at 95°C for 10min in a thermal cycler
(Boeco, Germany) before loading on a 10% SDS-PAGE gel.
Gels were afterwards stained with Coomassie blue [41], visu-
alized, and photographed using a gel documentation system
(UVP, Germany).
Murine pulmonary clearance model
The pulmonary clearance model in mice was carried out
as previously described [42]. Briefly, five groups (n = 6) of
6–8-week-old female BALB/C mice were infected intranasally
by injecting 40μl of a bacterial suspension (5 × 106CFU)
of each of the WT O35E, ΔfghA, ΔfghA/R, ΔadhC-fghA, and
ΔadhC-fghA/R into the nostrils under anesthesia using 250
µL of 25µg/mL 2,2,2-tribromoethanol. Four-and-a-half-hour
post-inoculation, mice were sacrificed by an overdose of the
anesthesia (750 µL of 25µg/mL 2,2,2-tribromoethanol), fol-
lowed by cervical dislocation. The lungs were excised, homog-
enized, serially diluted, and plated on TSB agar. Plates were
incubated for 48h followed by colony counting.
Results
M. catarrhalis possesses anEstD homolog
Using the N. meningitidis EstD as a query, Blastp against
the M. catarrhalis O35E proteome returned a 268 amino
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
Medical Microbiology and Immunology (2024) 213:3 Page 5 of 10 3
acid protein (EGE27440.1) annotated as FghA (for S-for-
mylglutathione hydrolase A) as the closest match, with a
query coverage of 96%, and an identity of 54% (Fig.S1).
Results of the NCBI BlastP between FghA of M. catarrhalis
strain O35E (EGE27440.1) and other M. catarrhalis strains
(taxid 480) shows a high degree of conservation with a per-
cent identity range of 91.79–100% (Fig. S2). Alignment of
homologs from different microbial genera, and even higher
eukaryotes showed that all had high identity (40–65%) and
similarity (55–77.7%) with the FghA of M. catarrhalis (Fig.
S3 and TableS1). The highest similarity obtained was with
S. pneumoniae (77.7%). Interestingly, although phylogenetic
analysis showed that the S. pneumoniae FghA is still the
closest evolutionary relative to the M. catarrhalis protein,
those from closely related genera which scored the high-
est similarity such as Neisseria, and Haemophilus clustered
together in a more distant branch of the tree. Meanwhile, the
FghA homolog from Paraccoccus showed a closer relation-
ship to the M. catarrhalis FghA, although it scored less on
sequence similarity than the aforementioned species (Fig.
S4).
We noticed that the ORF upstream of the M. catarrha-
lis fghA (Fig.1A) was annotated as adhC, a gene encod-
ing for S-hydroxymethyl alcohol dehydrogenase, and which
usually forms an operon with fghA as previously reported
[14, 20, 43]. To investigate the relationship between the
two ORFs, first, the operon prediction tool Operon Map-
per was used, and the results obtained demonstrated that
adhC and fghA are expected to form an operon by a high
probability of 0.97. Next, the protein interaction network
database, STRING, was used to investigate if a predicted
functional link between the two proteins is likely. The gen-
erated interaction network indicated that the M. catarrhalis
AdhC and FghA are strongly predicted as functional partners
by a score of 0.999. The two genes, Adhc and FghA, interact
by gene neighborhood, gene fusion, gene co-occurrence, co-
expression, and protein homology (Fig.1B). The network
also revealed the interaction by gene fusion between FghA
with EA1_06621, a Gdsl-like lipase/ccyl hydrolase family
protein. Another interaction by gene neighborhood between
FghA and MsrAB (EA1_03390), a trifunctional thioredoxin/
methionine sulfoxide reductase a/b protein was shown by the
network. The final interaction by gene neighborhood and
co-expression was predicted with GdhA (EA1_02207) an
NADP-specific glutamate dehydrogenase and FghA. It is
worth mentioning that while the confidence score of these
former interactions was relatively low (0.4), the GdhA was
the only protein predicted to interact with both FghA, and
AdhC. Although it lies upstream of adhC, it was not pre-
dicted to form an operon with the pair.
A similar blast analysis to that conducted using the
FghA was done using the M. catarrhalis O35E AdhC pro-
tein sequence. There was a high identity (41–81.7%) and
similarity (49.2%-90.9%) between AdhC of M. catarrhalis
and its homologs in other species (Fig. S5 and TableS2).
S. pneumoniae AdhC also showed the highest similarity
(90.9%) with that of M. catarrhalis, which is also in agree-
ment with the phylogenetic relationship (Fig. S6). As seen
with FghA, the AdhC proteins of closely related species
diverged from that of M. catarrhalis.
The putative adhC–fghA operon contributes
toformaldehyde detoxification inM. catarrhalis
To investigate the possible role of adhC and fghA in formal-
dehyde detoxification in M. catarrhalis, deletion mutants
lacking the fghA gene (ΔfghA) and both genes fghA and
adhC (ΔadhC-fghA) were constructed, and along with their
complemented mutants (ΔfghA/R and ΔadhC-fghA/R) were
tested for their formaldehyde sensitivity compared to WT
O35E using a disc diffusion susceptibility assay. There was
no significant difference between the five strains in sensitiv-
ity to formaldehyde using this susceptibility assay (Fig.2).
A more sensitive susceptibility assay was then performed
by plating serial dilutions of the five strains on a solid
medium containing increasing concentrations of formalde-
hyde (0, 0.8, and 1mM). Especially at the 1mM formalde-
hyde concentration, the growth of the ΔadhC-fghA was sig-
nificantly attenuated (Fig.3), while this growth defect was
not observed with the rescue strain ΔadhC-fghA/R which
reverted to the WT phenotype. Conversely, the single mutant
ΔfghA and its complemented mutant ΔfghA/R did not dis-
play significant sensitivity to formaldehyde at any of the
tested concentrations. It is worth mentioning here that all
constructed mutants showed no growth defects in compari-
son to O35E when the growth was monitored logarithmically
in plain TSB (Fig.S7).
Upon determination of the formaldehyde MIC for the five
strains, results very similar to those observed above were
obtained. For each of the four strains WT O35E, ΔfghA,
ΔfghA/R, and ΔadhC-fghA/R, the MIC value was 1mM.
On the other hand, ΔadhC-fghA exhibited much lower MIC
value of 7.8µM.
Loss oftheformaldehyde detoxification system
does notalter theprotein expression profile inM.
catarrhalis onSDS‑PAGE
To investigate whether the loss of AdhC and FghA could
alter the expression profile of some M. catarrhalis proteins,
especially given the results of the predicted interactions
obtained in silico from STRING which indicated the pres-
ence of a possible co-expression relationship between some
of the protein pairs, the profiles following incubation with
formaldehyde were visually assessed using SDS-PAGE. As
the only co-expression interactions noticed in silico were
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
Medical Microbiology and Immunology (2024) 213:3 3 Page 6 of 10
between FghA with AdhC and FghA with Gdha, no promi-
nent differences could be visually detected between the
WT and any of the ΔadhC-fghA and the ΔfghA mutants or
their complemented counterparts whether with or without
formaldehyde (Fig.S8). This indicates that any differences
that might exist are much more subtle to be detected using
Coomassie blue staining of SDS-PAGE gels.
The AdhC–FghA system isessential forpulmonary
colonization byM. catarrhalis
To investigate whether the role of the AdhC–FghA system
in formaldehyde detoxification would prove important to
the fitness of M. catarrhalis in pulmonary infection, an M.
catarrhalis pulmonary clearance model was conducted.
Interestingly, mice were significantly more capable of clear-
ing both ΔfghA and ΔadhC-fghA to a higher extent than WT
O35E and the complemented mutants with an average one-
log cycle change in the obtained colony counts (Fig.4).
These invivo results indicate that AdhC–FghA system sig-
nificantly contributes to the pathogenesis of M. catarrhalis.
Discussion
Formaldehyde is highly cytotoxic to living organisms, so
they need systems to detoxify formaldehyde to be able to
survive. Several researchers investigated the formaldehyde
detoxification mechanisms and the proteins involved to cope
with this stress [4, 6, 7, 16, 18, 20]. However, the current
study investigates the role of AdhC and FghA proteins in
formaldehyde detoxification in M. catarrhalis.
M. catarrhalis,previously known as N. catarrhalis,
resembles commensalNeisseriaspp. in culture, phenotype,
and ecological niche [44]. Therefore, Neisseria spp. previ-
ously characterized esterase protein EstD seemed to be a
good template to search for similar formaldehyde detoxi-
fying protein in M. catarrhalis. Our results indicated the
high conservation of AdhC and FghA proteins across M.
catarrhalis strains in addition to the species surveyed in
silico in this study. These results could be attributed to the
importance of the system for stress tolerance [4, 7, 20].
This was especially true for S. pneumoniae, which shared
the highest similarity and evolutionary relationship with the
M. catarrhalis proteins. The striking resemblance between
Fig. 2 Loss of the adhC–fghA
operon or fghA gene did not
impair formaldehyde detoxifi-
cation in M. catarrhalis when
tested by disc diffusion assay.
A Photographs of the zones
of inhibition were taken using
a gel documentation system
(UVP) with a ruler under each
strain that represents the scale
in cm. B Box plot representing
results of the formaldehyde disc
diffusion assay, plotting the
zone diameter obtained in cm.
The data represent the mean of
three independent experiments,
and the bars span the difference
between the minimum and max-
imum readings. The line inside
the box represents the median.
The graph was generated using
GraphPad prism version 9.0.0
(GraphPad Software, San
Diego, CA, USA). Statistical
analysis of the data was done
applying one-way ANOVA fol-
lowed by Tukey’s multiple com-
parison test. The “ns” stands for
“non-significant”
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
Medical Microbiology and Immunology (2024) 213:3 Page 7 of 10 3
these two organisms, which are considered two major causes
of acute otitis media, could be due to their shared habitat.
They share the same environment within the host, and this
could have driven the development of comparable protective
strategies to combat the same stresses encountered in the
nasopharynx and upper respiratory tract [44].
The interaction network of AdhC and FghA with each
other showed, most importantly, gene neighborhood, co-
expression, and protein homology. This gave plausible evi-
dence supporting the potential roles assigned to these pro-
teins, and their functional relationship. The network also
revealed the interaction by gene fusion between FghA with
other proteins like that encoded by EA1_06621, a Gdsl-like
lipase/ccyl hydrolase family protein. In addition, the MsrAB
(EA1_03390), a trifunctional thioredoxin/methionine sul-
foxide reductase a/b protein that has an important role as
a repair enzyme for proteins that have been inactivated by
oxidation [45] was predicted to have a gene neighborhood
interaction with the FghA. The MsrAB mode of function is
closely related to the mechanism of formaldehyde detoxifi-
cation through the redox potential of glutathione. Another
notable interaction was predicted with GdhA (EA1_02207),
a NADP-specific glutamate dehydrogenase, that belongs to
the Glu/Leu/Phe/Val dehydrogenases family. Glutamate
metabolism plays an essential role in the synthesis of glu-
tathione and Gdh-null mutants generally show a higher
sensitivity to oxidative stress as well as a more rapid deple-
tion of glutathione. These functions support the involvement
of GdhA in the thiol-dependent pathway of formaldehyde
detoxification [46, 47].
As expected, the Operon Mapper data indicated that adhC
and fghA very likely constitute an operon. These findings
are consistent with previous reports that also point out that
these two genes form an operon in other species [4, 7, 14,
20, 43]. As per previous bioinformatic analysis, an attempt
to confirm the function of adhC and fghA in formaldehyde
detoxification in M. catarrhalis was designed. At first, a
single mutant ΔfghA was constructed. There was no signifi-
cant difference between the wild-type and the single mutant
ΔfghA in the formaldehyde sensitivity tests, so a double
mutant ΔadhC-fghA was constructed. A study conducted by
Chen and co-workers mentioned that their single mutant of
the fghA homolog was more sensitive to formaldehyde than
their double mutant [20]. On the contrary, the current study
shows that the increase in sensitivity is more significant
in the double mutant ΔadhC-fghA than the single mutant
ΔfghA. The high sensitivity to formaldehyde in the double
mutant was expected as both proteins interact together in
the formaldehyde detoxification pathway, while in the single
mutant, the results showed that FghA is not essential on its
own to detoxify formaldehyde. One speculation could be
that, in the adhC–fghA mutant, due to the inactivation of
Fig. 3 Loss of the adhC–fghA
operon but not fghA impairs
formaldehyde detoxification
in M. catarrhalis when tested
by serial dilution susceptibility
assay. A Photographs of dilu-
tions on TSA plates containing
increasing concentrations of for-
maldehyde (0mM, 0.8mM, and
1mM) were taken using a gel
documentation system (UVP).
B Box plot graphs of bacterial
counts in CFU/mL with the
different formaldehyde con-
centrations. The data represent
the mean of three independent
experiments, and the bars span
the difference between the mini-
mum and maximum readings.
The line inside the box repre-
sents the median. The graphs
were generated using GraphPad
prism version 9.0.0 (GraphPad
Software, San Diego, California
USA). Statistical analysis of the
data was done applying one-way
ANOVA followed by Tukey’s
multiple comparison test
(*P ≤ 0.05), (**P ≤ 0.01). The
* indicates that the difference is
statistically significant
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
Medical Microbiology and Immunology (2024) 213:3 3 Page 8 of 10
theadhC, the accumulation of S-hydroxymethyl glutathione,
the substrate of AdhC, is more toxic to the cells than the
accumulation of S-formyl glutathione, the substrate of FghA,
in the fghA mutant [4]. A complementary explanation for
these results which would account for the non-toxicity of
S-formyl glutathione could be attributed to the presence of
another enzyme that participates with FghA, or completely
replaces it, in the hydrolysis of S-formylglutathione to glu-
tathione and formate. This explanation was also suggested
by Harms and co-workers where they found that fghA mutant
in Paracoccus denitrificans was able to grow on choline, a
formaldehyde-generating substrate, which comes in agree-
ment with our results [6]. Interestingly, during our bioinfor-
matic analysis, we found that the two E. coli proteins YeiG
and FrmB are two homologs for FghA of M. catarrhalis
O35E. Gonzalez and co-workers mentioned that the simul-
taneous deletion of both yeiG and frmB genes is required
to increase the sensitivity to formaldehyde, since the two
proteins contribute to the detoxification of formaldehyde [7].
This could be the case in M. catarrhalis, with a structural
homolog to FghA rather than a sequence homolog that is yet
to be discovered, as none could be found in the proteome
of M. catarrhalis using blast analysis. A study conducted
by Potter and co-workers reported that each of the single
mutants ΔestD and ΔadhC showed identical zones of inhi-
bition to that of the WT by the disc diffusion formaldehyde
sensitivity assay [17]. In the current work, the disc diffusion
method could not differentiate the susceptibilities of the con-
structed mutants and their wild-type counterpart. However,
upon using more quantitative methods, the differences in
the formaldehyde susceptibilities were more prominent. It
was found that a significant difference exists between the
adhC–fghA mutant tested compared to the wild type which
demonstrated that the system contributes to formaldehyde
detoxification.
Previous studies have revealed the important role played
by the adhC–fghA system in bacterial virulence [17, 20].
In the current work, both the ΔfghA and the ΔadhC-fghA
mutants show marked decreased fitness in a pulmonary
clearance model as is evidenced by the significantly lower
colony counts retrieved from mice infected with these
strains. This shows that regardless of its direct role in for-
maldehyde detoxification, FghA could contribute to the
resistance to clearance of M. catarrhalis from the host cells
by a mechanism that is yet to be elucidated. Interestingly,
the obtained data come in accordance with a study which
showed that an FghA homolog, EstD in N. gonorrhoeae, is
necessary for bacterial growth in the host’s cervical epithe-
lial cells, although it did not show a formaldehyde-sensitive
phenotype using disc diffusion assay. The mentioned study
reported that EstD had a potential role in the nitrosative
stress defense system ofN. gonorrhoeae which allows it to
counteract the killing effect of nitric oxide [9] released by
phagocytic cells in the inflammatory response to infection
[17]. This shows that FghA could be involved in combat-
ing other kinds of stresses; hence, its role in pathogenesis
warrants future studies. The significant increase in the pul-
monary clearance extent of ΔfghA as well as ΔadhC-fghA
compared to WT O35E reveals the necessity of both genes
for survival in the respiratory tract of the host. Moreover, the
current findings are similar to a previous study that pointed
out to the possible important role the AdhC–EstD system
plays in the survival and virulence of N. meningitidis, after
observing that the ΔadhC–estD mutant is non-viable in
experimentally induced biofilms [20].
To put this in a more biologically relevant context, Chen
and co-workers reported that the concentration of formal-
dehyde in the blood of healthy individuals is estimated to
be around 0.1mM [4]. However, it is assumed that dur-
ing inflammation, infection, and the associated respiratory
burst, the localized concentration of formaldehyde would
rise above the normal non-toxic levels. Furthermore, the
Fig. 4 Loss of the adhC–fghA operon and fghA results in increased
pulmonary clearance of M. catarrhalis in the murine pulmo-
nary infection model. A box plot representing the colony counts
in log10 CFU/mL, of M. catarrhalis WT O35E, ΔfghA, ΔfghA/R,
ΔadhC-fghA double mutant, and ΔadhC-fghA/R as obtained from
the lungs of infected mice. The bars span the difference between the
minimum and maximum readings. The + sign represents the mean of
the log10CFU/mL. The graphs were generated using GraphPad prism
version 9.0.0 (GraphPad Software, San Diego, CA, USA). Statisti-
cal analysis of data was performed by one-way ANOVA in GrapPad
Prism. The * indicates that the difference is statistically significant
as determined by Tukey’s multiple comparison test (**P ≤ 0.01),
(***P ≤ 0.001), and (****P ≤ 0.0001)
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
Medical Microbiology and Immunology (2024) 213:3 Page 9 of 10 3
complexity of the host response on the molecular level
would certainly mean that other stresses are present and
could contribute to the significantly higher clearance
observed with the mutants.
Conclusion
This study reports for the first time anfghA mutant and an
adhC/fghA double mutantphenotypes in the emerging path-
ogenM. catarrhalis.The findings in this research indicate
that the system plays a crucial role in formaldehyde detoxifi-
cation and in lung colonization by M. catarrhalis. Moreover,
these findings shed light on the importance of understanding
the adhC–fghA system inM. catarrhalisto help in the poten-
tial development of novel therapeutics to combat infections
caused by this emerging drug-resistant pathogen.
Supplementary Information The online version contains supplemen-
tary material available at https:// doi. org/ 10. 1007/ s00430- 024- 00785-0.
Author contributions All authors have contributed to the conceptu-
alization of the idea. D.O, N.M.E, and A.S.A have contributed to the
methodology and data curation. All authors have contributed to the
analysis and interpretation of the data. D.O and N.M.E. have contrib-
uted to the preparation of the original draft: All authors have read,
revised, and approved the final version of the manuscript.
Funding Open access funding provided by The Science, Technology &
Innovation Funding Authority (STDF) in cooperation with The Egyp-
tian Knowledge Bank (EKB). No funding was received for conducting
this study.
Declarations
Conflict of interest The authors have no relevant financial or non-fi-
nancial interests to disclose.
Open Access This article is licensed under a Creative Commons Attri-
bution 4.0 International License, which permits use, sharing, adapta-
tion, distribution and reproduction in any medium or format, as long
as you give appropriate credit to the original author(s) and the source,
provide a link to the Creative Commons licence, and indicate if changes
were made. The images or other third party material in this article are
included in the article’s Creative Commons licence, unless indicated
otherwise in a credit line to the material. If material is not included in
the article’s Creative Commons licence and your intended use is not
permitted by statutory regulation or exceeds the permitted use, you will
need to obtain permission directly from the copyright holder. To view a
copy of this licence, visit http://creativecommons.org/licenses/by/4.0/.
References
1. Sillanpää S etal (2016) Moraxella catarrhalis might be more
common than expected in acute otitis media in young Finnish
children. J Clin Microbiol 54(9):2373–2379
2. Gupta DN, Arora S, Kundra S (2011) Moraxella catarrhalis as a
respiratory pathogen. Indian J Pathol Microbiol 54:769–771
3. Raveendran S etal (2020) Moraxella catarrhalis: a cause of con-
cern with emerging resistance and presence of bro beta-lactamase
gene—report from a Tertiary Care Hospital in South India. Inter-
national Journal of Microbiology 2020:7316257
4. Chen NH etal (2016) Formaldehyde stress responses in bacterial
pathogens. Front Microbiol 7:257
5. Klein VJ etal (2022) Unravelling formaldehyde metabolism in
bacteria: road towards synthetic methylotrophy. Microorganisms
10(2):220
6. Harms N et al (1996) S-formylglutathione hydrolase of Para-
coccus denitrificans is homologous to human esterase D: a
universal pathway for formaldehyde detoxification? J Bacteriol
178(21):6296–6299
7. Gonzalez CF et al (2006) Molecular basis of formaldehyde
detoxification. Characterization of two S-formylglutathione
hydrolases from Escherichia coli, FrmB and YeiG. J Biol Chem
281(20):14514–14522
8. Mason, R.P., etal., Formaldehyde metabolism by Escherichia coli.
Detection by invivo 13C NMR spectroscopy of S-(hydroxymethyl)
glutathione as a transient intracellular intermediate. Biochemis-
try, 1986. 25(16): p. 4504–7.
9. Wilson SM, Gleisten MP, Donohue TJ (2008) Identification of
proteins involved in formaldehyde metabolism by Rhodobacter
sphaeroides. Microbiology (Reading) 154(Pt 1):296–305
10. Goenrich M etal (2002) A glutathione-dependent formaldehyde-
activating enzyme (Gfa) from Paracoccus denitrificans detected
and purified via two-dimensional proton exchange NMR spectros-
copy. J Biol Chem 277(5):3069–3072
11. Uotila L, Koivusalo M (1974) Formaldehyde dehydrogenase from
human liver. Purification, properties, and evidence for the for-
mation of glutathione thiol esters by the enzyme. J Biol Chem
249(23):7653–7663
12. Uotila L, Koivusalo M (1974) Purification and properties of
S-formylglutathione hydrolase from human liver. J Biol Chem
249(23):7664–7672
13. Hopkinson RJ etal (2015) Studies on the glutathione-dependent
formaldehyde-activating enzyme from Paracoccus denitrificans.
PLoS ONE 10(12):e0145085
14. Kidd SP etal (2007) Glutathione-dependent alcohol dehydroge-
nase AdhC is required for defense against nitrosative stress in
Haemophilus influenzae. Infect Immun 75(9):4506–4513
15. Wongsaroj L etal (2018) Pseudomonas aeruginosa glutathione
biosynthesis genes play multiple roles in stress protection, bacte-
rial virulence and biofilm formation. PLoS ONE 13(10):e0205815
16. Jamet A etal (2013) Identification of genes involved in Neisseria
meningitidis colonization. Infect Immun 81(9):3375–3381
17. Potter AJ etal (2009) Esterase D is essential for protection of
Neisseria gonorrhoeae against nitrosative stress and for bacterial
growth during interaction with cervical epithelial cells. J Infect
Dis 200(2):273–278
18. Potter AJ etal (2010) The MerR/NmlR family transcription fac-
tor of Streptococcus pneumoniae responds to carbonyl stress
and modulates hydrogen peroxide production. J Bacteriol
192(15):4063–4066
19. Yurimoto H etal (2003) Physiological role of S-formylglutathione
hydrolase in C(1) metabolism of the methylotrophic yeast Can-
dida boidinii. Microbiology (Reading) 149(Pt 8):1971–1979
20. Chen NH etal (2013) A glutathione-dependent detoxification
system is required for formaldehyde resistance and optimal sur-
vival of Neisseria meningitidis in biofilms. Antioxid Redox Signal
18(7):743–755
21. Hoopman TC etal (2011) Identification of gene products involved
in the oxidative stress response of Moraxella catarrhalis. Infect
Immun 79(2):745–755
22. Spaniol V, Bernhard S, Aebi C (2015) Moraxella catarrhalis
AcrAB-OprM efflux pump contributes to antimicrobial resistance
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
Medical Microbiology and Immunology (2024) 213:3 3 Page 10 of 10
and is enhanced during cold shock response. Antimicrob Agents
Chemother 59(4):1886–1894
23. Zeng Q etal (2020) A moraxella virulence factor catalyzes an
essential esterase reaction of biotin biosynthesis. Front Microbiol
11:148
24. Unhanand M et al (1992) Pulmonary clearance of Moraxella
catarrhalis in an animal model. J Infect Dis 165(4):644–650
25. Lewis J, Gorman G, Chaney T (1974) Methods to determine CO2
levels in a gonococcal transport system (Transgrow). Health Lab
Sci 11(2):65–68
26. Altschul SF etal (1990) Basic local alignment search tool. J Mol
Biol 215(3):403–410
27. Sievers F, Higgins DG (2018) Clustal Omega for making accurate
alignments of many protein sequences. Protein Sci 27(1):135–145
28. Junier T, Zdobnov EM (2010) The Newick utilities: high-through-
put phylogenetic tree processing in the UNIX shell. Bioinformat-
ics 26(13):1669–1670
29. Katoh K, Standley DM (2013) MAFFT multiple sequence align-
ment software version 7: improvements in performance and usa-
bility. Mol Biol Evol 30(4):772–780
30. Criscuolo A, Gribaldo S (2010) BMGE (block mapping and gath-
ering with entropy): a new software for selection of phylogenetic
informative regions from multiple sequence alignments. BMC
Evol Biol 10(1):210
31. Guindon S etal (2010) New algorithms and methods to estimate
maximum-likelihood phylogenies: assessing the performance of
PhyML 3.0. Syst Biol 59(3):307–321
32. Lemoine F etal (2018) Renewing Felsenstein’s phylogenetic boot-
strap in the era of big data. Nature 556(7702):452–456
33. McWilliam H etal (2013) Analysis tool web services from the
EMBL-EBI. Nucleic Acids Res 41(Web Server issue):W597-600
34. Szklarczyk D et al (2019) STRING v11: protein-protein asso-
ciation networks with increased coverage, supporting functional
discovery in genome-wide experimental datasets. Nucleic Acids
Res 47(D1):D607-d613
35. Taboada B etal (2018) Operon-mapper: a web server for precise
operon identification in bacterial and archaeal genomes. Bioinfor-
matics 34(23):4118–4120
36. Menard R, Sansonetti PJ, Parsot C (1993) Nonpolar mutagen-
esis of the Ipa genes defines IpaB, IpaC, and IpaD as effec-
tors of Shigella flexneri entry into epithelial cells. J Bacteriol
175(18):5899–5906
37. Pearson MM etal (2006) Biofilm formation by Moraxella catarrh-
alis invitro: roles of the UspA1 adhesin and the Hag hemagglu-
tinin. Infect Immun 74(3):1588–1596
38. Attia AS etal (2005) The UspA2 protein of Moraxella catarrhalis
is directly involved in the expression of serum resistance. Infect
Immun 73(4):2400–2410
39. Nair J etal (1993) The rpsL gene and streptomycin resistance
in single and multiple drug-resistant strains of Mycobacterium
tuberculosis. Mol Microbiol 10(3):521–527
40. CLSI (2018) Reference method for broth microdilution antibacte-
rial susceptibility testing; approved standard-11th edition, in CLSI
document M07–A11. Clinical and Laboratory Standards Institute,
Wayne
41. Russell JSADW (2001) Molecular cloning: a laboratory manual,
3rd edn. Cold Spring Harbor Laboratory Press, New York
42. Yassin GM, Amin MA, Attia AS (2016) Immunoinformatics
identifies a lactoferrin binding protein a peptide as a promising
vaccine with a global protective prospective against Moraxella
catarrhalis. J Infect Dis 213(12):1938–1945
43. Herring CD, Blattner FR (2004) Global transcriptional effects of
a suppressor tRNA and the inactivation of the regulator frmR. J
Bacteriol 186(20):6714–6720
44. Goldstein EJC, Murphy TF, Parameswaran GI (2009) Moraxella
catarrhalis, a human respiratory tract pathogen. Clin Infect Dis
49(1):124–131
45. Jen FE-C etal (2019) The Neisseria gonorrhoeae methionine sul-
foxide reductase (MsrA/B) is a surface exposed, immunogenic,
vaccine candidate. Front Immunol 10:137
46. Osman D etal (2016) The effectors and sensory sites of formal-
dehyde-responsive regulator FrmR and metal-sensing variant. J
Biol Chem 291(37):19502–19516
47. Lee YJ etal (2012) Involvement of GDH3-encoded NADP+-
dependent glutamate dehydrogenase in yeast cell resistance to
stress-induced apoptosis in stationary phase cells. J Biol Chem
287(53):44221–44233
Publisher's Note Springer Nature remains neutral with regard to
jurisdictional claims in published maps and institutional affiliations.
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
1.
2.
3.
4.
5.
6.
Terms and Conditions
Springer Nature journal content, brought to you courtesy of Springer Nature Customer Service Center GmbH (“Springer Nature”).
Springer Nature supports a reasonable amount of sharing of research papers by authors, subscribers and authorised users (“Users”), for small-
scale personal, non-commercial use provided that all copyright, trade and service marks and other proprietary notices are maintained. By
accessing, sharing, receiving or otherwise using the Springer Nature journal content you agree to these terms of use (“Terms”). For these
purposes, Springer Nature considers academic use (by researchers and students) to be non-commercial.
These Terms are supplementary and will apply in addition to any applicable website terms and conditions, a relevant site licence or a personal
subscription. These Terms will prevail over any conflict or ambiguity with regards to the relevant terms, a site licence or a personal subscription
(to the extent of the conflict or ambiguity only). For Creative Commons-licensed articles, the terms of the Creative Commons license used will
apply.
We collect and use personal data to provide access to the Springer Nature journal content. We may also use these personal data internally within
ResearchGate and Springer Nature and as agreed share it, in an anonymised way, for purposes of tracking, analysis and reporting. We will not
otherwise disclose your personal data outside the ResearchGate or the Springer Nature group of companies unless we have your permission as
detailed in the Privacy Policy.
While Users may use the Springer Nature journal content for small scale, personal non-commercial use, it is important to note that Users may
not:
use such content for the purpose of providing other users with access on a regular or large scale basis or as a means to circumvent access
control;
use such content where to do so would be considered a criminal or statutory offence in any jurisdiction, or gives rise to civil liability, or is
otherwise unlawful;
falsely or misleadingly imply or suggest endorsement, approval , sponsorship, or association unless explicitly agreed to by Springer Nature in
writing;
use bots or other automated methods to access the content or redirect messages
override any security feature or exclusionary protocol; or
share the content in order to create substitute for Springer Nature products or services or a systematic database of Springer Nature journal
content.
In line with the restriction against commercial use, Springer Nature does not permit the creation of a product or service that creates revenue,
royalties, rent or income from our content or its inclusion as part of a paid for service or for other commercial gain. Springer Nature journal
content cannot be used for inter-library loans and librarians may not upload Springer Nature journal content on a large scale into their, or any
other, institutional repository.
These terms of use are reviewed regularly and may be amended at any time. Springer Nature is not obligated to publish any information or
content on this website and may remove it or features or functionality at our sole discretion, at any time with or without notice. Springer Nature
may revoke this licence to you at any time and remove access to any copies of the Springer Nature journal content which have been saved.
To the fullest extent permitted by law, Springer Nature makes no warranties, representations or guarantees to Users, either express or implied
with respect to the Springer nature journal content and all parties disclaim and waive any implied warranties or warranties imposed by law,
including merchantability or fitness for any particular purpose.
Please note that these rights do not automatically extend to content, data or other material published by Springer Nature that may be licensed
from third parties.
If you would like to use or distribute our Springer Nature journal content to a wider audience or on a regular basis or in any other manner not
expressly permitted by these Terms, please contact Springer Nature at
onlineservice@springernature.com
ResearchGate has not been able to resolve any citations for this publication.
Article
Full-text available
Formaldehyde metabolism is prevalent in all organisms, where the accumulation of formaldehyde can be prevented through the activity of dissimilation pathways. Furthermore, formaldehyde assimilatory pathways play a fundamental role in many methylotrophs, which are microorganisms able to build biomass and obtain energy from single- and multicarbon compounds with no carbon-carbon bonds. Here, we describe how formaldehyde is formed in the environment, the mechanisms of its toxicity to the cells, and the cell's strategies to circumvent it. While their importance is unquestionable for cell survival in formaldehyde rich environments, we present examples of how the modification of native formaldehyde dissimilation pathways in nonmethylotrophic bacteria can be applied to redirect carbon flux toward heterologous, synthetic formaldehyde assimilation pathways introduced into their metabolism. Attempts to engineer methylotrophy into nonmethylotrophic hosts have gained interest in the past decade, with only limited successes leading to the creation of autonomous synthetic methylotrophy. Here, we discuss how native formaldehyde assimilation pathways can additionally be employed as a premise to achieving synthetic methylotrophy. Lastly, we discuss how emerging knowledge on regulation of formaldehyde metabolism can contribute to creating synthetic regulatory circuits applied in metabolic engineering strategies.
Article
Full-text available
We have purified formaldehyde dehydrogenase (EC 1.2.1.1) 1390-fold from human liver. The final preparation, which has a specific activity of 3.20 i.u. per mg of protein (25°) is homogeneous according to electrophoretic criteria. S-Formylglutathione rather than formate is formed from formaldehyde and reduced glutathione in the reaction catalyzed by purified formaldehyde dehydrogenase. The enzyme is not strictly NAD-specific; NADP can also be used although NADP has a much higher Km value than NAD, especially at high pH values. S-Formylglutathione is reduced by the enzyme to formaldehyde with either NADH or NADPH as cofactors. Methylglyoxal and some other ketoaldehydes also can be used as the substrates of formaldehyde dehydrogenase. The product obtained from methylglyoxal is probably S-pyruvylglutathione. S-Formylglutathione is hydrolyzed very actively in crude human liver preparations. All of this activity is removed during the purification of formaldehyde dehydrogenase. Formaldehyde dehydrogenase has an apparent molecular weight of 81,400 according to gel filtration and a subunit molecular weight of 39,500 according to dodecyl sulfate gel electrophoresis. Isoelectric focusing experiments gave an isoelectric point, pI of 6.35. The enzyme is very sensitive to mercaptide-forming reagents. NAD and NADH protect the enzyme. NAD and especially NADH stabilize the enzyme during storage and against denaturation at high temperatures.
Article
Full-text available
Pimeloyl-acyl carrier protein (ACP) methyl ester esterase catalyzes the last biosynthetic step of the pimelate moiety of biotin, a key intermediate in biotin biosynthesis. The paradigm pimeloyl-ACP methyl ester esterase is the BioH protein of Escherichia coli that hydrolyses the ester bond of pimeloyl-ACP methyl ester. Biotin synthesis in E. coli also requires the function of the malonyl-ACP methyltransferase gene (bioC) to employ a methylation strategy to allow elongation of a temporarily disguised malonate moiety to a pimelate moiety by the fatty acid synthetic enzymes. However, bioinformatics analyses of the extant bacterial genomes showed that bioH is absent in many bioC-containing bacteria. The genome of the Gram-negative bacterium, Moraxella catarrhalis lacks a gene encoding a homolog of any of the six known pimeloyl-ACP methyl ester esterase isozymes suggesting that this organism encodes a novel pimeloyl-ACP methyl ester esterase isoform. We report that this is the case. The gene encoding the new isoform, called btsA, was isolated by complementation of an E. coli bioH deletion strain. The requirement of BtsA for the biotin biosynthesis in M. catarrhalis was confirmed by a biotin auxotrophic phenotype caused by deletion of btsA in vivo and a reconstituted in vitro desthiobiotin synthesis system. Purified BtsA was shown to cleave the physiological substrate pimeloyl-ACP methyl ester to pimeloyl-ACP by use of a Ser117-His254-Asp287 catalytic triad. The lack of sequence alignment with other isozymes together with phylogenetic analyses revealed BtsA as a new class of pimeloyl-ACP methyl ester esterase. The involvement of BtsA in M. catarrhalis virulence was confirmed by the defect of bacterial invasion to lung epithelial cells and survival within macrophages in the ΔbtsA strains. Identification of the new esterase gene btsA exclusive in Moraxella species that links biotin biosynthesis to bacterial virulence, can reveal a new valuable target for development of drugs against M. catarrhalis.
Article
Full-text available
Background. Found as a commensal in the upper respiratory tract, Gram-negative diplococcus Moraxella catarrhalis did not hold much importance as an infectious agent for long. The emergence of the first antibiotic-resistant strain of M. catarrhalis was noted in 1977 in Sweden. This has gradually spread worldwide over the years to more than 95% of the strains showing resistance to penicillin now. Penicillin resistance is mediated by the production of beta-lactamases encoded by bro-1 and bro-2 genes that code for beta-lactamases BRO-1 and BRO-2, respectively. The purpose of this study was to explore the trends of antibiotic resistance, the presence of bro genes, and clinical correlation of these findings with the rise in M. catarrhalis infections worldwide. Methods. Strains of M. catarrhalis were isolated from the respiratory samples submitted to the microbiology laboratory. Preliminary identification was done using standard microbiological techniques, and antibiotic sensitivity was determined by minimum inhibitory concentration assessed using the E-test. Further, the genes associated with the development of resistance to penicillin (beta-lactamase enzyme) were detected using polymerase chain reaction technique. Results. Fourteen strains of M. catarrhalis were isolated during the study period. Majority of the strains were isolated from patients between 40 and 60 years of age and from males. Seasonality was observed with most strains being isolated during the winter season. The most important predisposing factors identified were advanced age with a history of smoking and chronic obstructive pulmonary disease. The antibiotic susceptibility pattern showed resistance to most antibiotics commonly used for the treatment of respiratory tract infections. Finally, all the strains were beta-lactamase producers, confirmed by the detection of bro-1 beta-lactamase gene in them. Conclusion. The increase in antibiotic resistance and beta-lactamase production in M. catarrhalis is a cause of concern. The emerging resistance pattern emphasises the need for an appropriate antibiotic stewardship program in clinical practice. Importance should be given to the monitoring of the trends of antibiotic susceptibility and their usage to prevent the emergence of outbreaks with resistant strains and treatment failures. 1. Introduction WHO has recognised antibiotic resistance as a global threat. What were earlier considered as harmless infections have now become very difficult to treat because they are caused by organisms that have developed resistance to commonly used antibiotics. One such organism, in which the rise in antibiotic resistance has been alarmingly high, is Moraxella catarrhalis. M. catarrhalis is a Gram-negative, aerobic, oxidase-positive diplococcus frequently seen as a coloniser of the upper respiratory tract [1]. The organism, in its pathogenic state, causes upper respiratory tract infections such as otitis media in children and lower respiratory tract infection in adults [1]. The severity of infection usually depends on the host’s immune status. After the identification of beta-lactamase-producing M. catarrhalis strains in 1977 in Sweden, there has been an unprecedented increase in the beta-lactam antibiotic resistance in this organism throughout the world [2]. More than 95% of global clinical isolates are now resistant to penicillin [3–6], and resistance to other classes of antibiotics is also on the rise. This might have led M. catarrhalis to become a well-established pathogen rather than an emerging one. Beta-lactamases produced by the M. catarrhalis not only protect the pathogen but also inactivate penicillin, an antibiotic that is commonly used for the treatment of mixed infections caused by other airway pathogens such as Streptococcus pneumoniae and/or nontypeable Haemophilus influenzae [1, 4]. Molecular investigations reveal the presence of two types of beta-lactamases in M. catarrhalis, BRO-1 and BRO-2 that are encoded by two genes bro-1 and bro-2, respectively. These two enzymes can be distinguished by the presence of a 21-base pair (bp) deletion in the promoter region of the bro-2 gene when comparing the same region in the bro-1 gene [7, 8]. 2. Materials and Methods M. catarrhalis was isolated from respiratory clinical samples such as sputum, tracheal aspirates, and bronchoalveolar lavage (BAL) collected from December 2012 to December 2013 in the Microbiology Department, St. John’s Medical College, Bangalore. Clinical data for the patients from whom these strains were isolated were collected retrospectively. Ethical clearance for the study was obtained from the Institutional Ethics Committee (IEC) of St. John’s Medical College (IEC study Ref no: 201/2018). 2.1. Microbiological Methods Samples (sputum, tracheal trap, and BAL) were subjected to Gram staining (Figure 1) and then plated on Blood Agar (BA) and Chocolate Agar (CA) with bacitracin disc and MacConkey Agar (MA), following the standard operating procedure of the laboratory. Identification of the isolates was done by following the standard microbiological techniques.
Article
Full-text available
Control of the sexually transmitted infection gonorrhea is a major public health challenge, due to the recent emergence of multidrug resistant strains of Neisseria gonorrhoeae, and there is an urgent need for novel therapies or a vaccine to prevent gonococcal disease. In this study, we evaluated the methionine sulfoxide reductase (MsrA/B) of N. gonorrhoeae as a potential vaccine candidate, in terms of its expression, sequence conservation, localization, immunogenicity, and the functional activity of antibodies raised to it. Gonococcal MsrA/B has previously been shown to reduce methionine sulfoxide [Met(O)] to methionine (Met) in oxidized proteins and protect against oxidative stress. Here we have shown that the gene encoding MsrA/B is present, highly conserved, and expressed in all N. gonorrhoeae strains investigated, and we determined that MsrA/B is surface is exposed on N. gonorrhoeae. Recombinant MsrA/B is immunogenic, and mice immunized with MsrA/B and either aluminum hydroxide gel adjuvant or Freund's adjuvant generated a humoral immune response, with predominantly IgG1 antibodies. Higher titers of IgG2a, IgG2b, and IgG3 were detected in mice immunized with MsrA/B-Freund's adjuvant compared to MsrA/B-aluminum hydroxide adjuvant, while IgM titers were similar for both adjuvants. Antibodies generated by MsrA/B-Freund's in mice mediated bacterial killing via both serum bactericidal activity and opsonophagocytic activity. Anti-MsrA/B was also able to functionally block the activity of MsrA/B by inhibiting binding to its substrate, Met(O). We propose that recombinant MsrA/B is a promising vaccine antigen for N. gonorrhoeae.
Article
Full-text available
Proteins and their functional interactions form the backbone of the cellular machinery. Their connectivity network needs to be considered for the full understanding of biological phenomena, but the available information on protein-protein associations is incomplete and exhibits varying levels of annotation granularity and reliability. The STRING database aims to collect, score and integrate all publicly available sources of protein-protein interaction information, and to complement these with computational predictions. Its goal is to achieve a comprehensive and objective global network, including direct (physical) as well as indirect (functional) interactions. The latest version of STRING (11.0) more than doubles the number of organisms it covers, to 5090. The most important new feature is an option to upload entire, genome-wide datasets as input, allowing users to visualize subsets as interaction networks and to perform gene-set enrichment analysis on the entire input. For the enrichment analysis, STRING implements well-known classification systems such as Gene Ontology and KEGG, but also offers additional, new classification systems based on high-throughput text-mining as well as on a hierarchical clustering of the association network itself. The STRING resource is available online at https://string-db.org/.
Article
Full-text available
Pseudomonas aeruginosa PAO1 contains gshA and gshB genes, which encode enzymes involved in glutathione (GSH) biosynthesis. Challenging P. aeruginosa with hydrogen peroxide, cumene hydroperoxide, and t-butyl hydroperoxide increased the expression of gshA and gshB. The physiological roles of these genes in P. aeruginosa oxidative stress, bacterial virulence, and biofilm formation were examined using P. aeruginosa ΔgshA, ΔgshB, and double ΔgshAΔgshB mutant strains. These mutants exhibited significantly increased susceptibility to methyl viologen, thiol-depleting agent, and methylglyoxal compared to PAO1. Expression of functional gshA, gshB or exogenous supplementation with GSH complemented these phenotypes, which indicates that the observed mutant phenotypes arose from their inability to produce GSH. Virulence assays using a Drosophila melanogaster model revealed that the ΔgshA, ΔgshB and double ΔgshAΔgshB mutants exhibited attenuated virulence phenotypes. An analysis of virulence factors, including pyocyanin, pyoverdine, and cell motility (swimming and twitching), showed that these levels were reduced in these gsh mutants compared to PAO1. In contrast, biofilm formation increased in mutants. These data indicate that the GSH product and the genes responsible for GSH synthesis play multiple crucial roles in oxidative stress protection, bacterial virulence and biofilm formation in P. aeruginosa.
Article
Full-text available
Operon-mapper is a web server that accurately, easily, and directly predicts the operons of any bacterial or archaeal genome sequence. The operon predictions are based on the intergenic distance of neighboring genes as well as the functional relationships of their protein-coding products. To this end, Operon-mapper finds all the ORFs within a given nucleotide sequence, along with their genomic coordinates, orthology groups, and functional relationships. We believe that Operon-mapper, due to its accuracy, simplicity and speed, as well as the relevant information that it generates, will be a useful tool for annotating and characterizing genomic sequences. Availability: http://biocomputo.ibt.unam.mx/operon_mapper/.
Article
Full-text available
Felsenstein's application of the bootstrap method to evolutionary trees is one of the most cited scientific papers of all time. The bootstrap method, which is based on resampling and replications, is used extensively to assess the robustness of phylogenetic inferences. However, increasing numbers of sequences are now available for a wide variety of species, and phylogenies based on hundreds or thousands of taxa are becoming routine. With phylogenies of this size Felsenstein's bootstrap tends to yield very low supports, especially on deep branches. Here we propose a new version of the phylogenetic bootstrap in which the presence of inferred branches in replications is measured using a gradual 'transfer' distance rather than the binary presence or absence index used in Felsenstein's original version. The resulting supports are higher and do not induce falsely supported branches. The application of our method to large mammal, HIV and simulated datasets reveals their phylogenetic signals, whereas Felsenstein's bootstrap fails to do so.
Article
Full-text available
Clustal Omega is a widely used package for carrying out multiple sequence alignment. Here we describe some recent additions to the package and benchmark some alternative ways of making alignments. These benchmarks are based on protein structure comparisons or predictions and include a recently described method based on secondary structure prediction. In general Clustal Omega is fast enough to make very large alignments and the accuracy of protein alignments is high when compared to alternative packages. The package is freely available as executables or source code from www.clustal.org or can be run on-line from a variety of sites, especially the EBI www.ebi.ac.uk . This article is protected by copyright. All rights reserved.