ArticlePDF Available

The N Terminus of the PduB Protein Binds the Protein Shell of the Pdu Microcompartment to Its Enzymatic Core

American Society for Microbiology
Journal of Bacteriology
Authors:

Abstract and Figures

Bacterial microcompartments (MCPs) are extremely large proteinaceous organelles that consist of an enzymatic core encapsulated within a complex protein shell. A key question in MCP biology is the nature of the interactions that guide the assembly of thousands of protein subunits into a well-ordered metabolic compartment. In this report, we show that the N-terminal 37 amino acids of the PduB protein has a critical role in binding the shell of the 1,2-propanediol utilization (Pdu) microcompartment to its enzymatic core. Several mutations were constructed that deleted short regions of the N-terminus of PduB. Growth tests indicated that three of these deletions were impaired MCP assembly. Attempts to purify MCPs from these mutants, followed by gel electrophoresis and enzyme assays, indicated that the protein complexes isolated consisted of MCP shells depleted in core enzymes. Electron microscopy substantiated these findings by identifying apparently empty MCP shells but not intact MCPs. Analyses of 13 site-directed mutants indicated that the key region of the N-terminus of PduB required for MCP assembly is a putative helix spanning residues 6 to 18. Considering the findings presented here together with prior work, we propose a new model for MCP assembly. Importance Bacterial microcompartments consist of metabolic enzymes encapsulated within a protein shell and are widely used to optimize metabolic process. Here, we show that the N-terminal 37 amino acids of the PduB shell protein is essential for assembly of the 1,2-propanediol utilization microcompartment. Results indicate that it plays a key role in binding the outer shell to the enzymatic core. We propose that this interaction might be used to define the relative orientation of the shell with respect to the core. This finding is of fundamental importance to our understanding of microcompartment assembly and may have application to engineering microcompartments as nanobioreactors for chemical production.
This content is subject to copyright. Terms and conditions apply.
The N Terminus of the PduB Protein
Binds the Protein Shell of the Pdu
Microcompartment to Its Enzymatic Core
Brent P. Lehman, Chiranjit Chowdhury, Thomas A. Bobik
Roy J. Carver Department of Biochemistry, Biophysics and Molecular Biology, Iowa State University, Ames,
Iowa, USA
ABSTRACT Bacterial microcompartments (MCPs) are extremely large proteinaceous
organelles that consist of an enzymatic core encapsulated within a complex protein
shell. A key question in MCP biology is the nature of the interactions that guide the
assembly of thousands of protein subunits into a well-ordered metabolic compart-
ment. In this report, we show that the N-terminal 37 amino acids of the PduB pro-
tein have a critical role in binding the shell of the 1,2-propanediol utilization (Pdu)
microcompartment to its enzymatic core. Several mutations were constructed that
deleted short regions of the N terminus of PduB. Growth tests indicated that three
of these deletions were impaired MCP assembly. Attempts to purify MCPs from
these mutants, followed by gel electrophoresis and enzyme assays, indicated that
the protein complexes isolated consisted of MCP shells depleted of core enzymes.
Electron microscopy substantiated these findings by identifying apparently empty
MCP shells but not intact MCPs. Analyses of 13 site-directed mutants indicated that
the key region of the N terminus of PduB required for MCP assembly is a putative
helix spanning residues 6 to 18. Considering the findings presented here together
with prior work, we propose a new model for MCP assembly.
IMPORTANCE Bacterial microcompartments consist of metabolic enzymes encapsu-
lated within a protein shell and are widely used to optimize metabolic process. Here,
we show that the N-terminal 37 amino acids of the PduB shell protein are essential
for assembly of the 1,2-propanediol utilization microcompartment. The results indi-
cate that it plays a key role in binding the outer shell to the enzymatic core. We
propose that this interaction might be used to define the relative orientation of the
shell with respect to the core. This finding is of fundamental importance to our un-
derstanding of microcompartment assembly and may have application to engineer-
ing microcompartments as nanobioreactors for chemical production.
KEYWORDS microcompartment, carboxysome, 1,2-propanediol, Salmonella,
vitamin B
12
Hundreds of species of bacteria produce complex proteinaceous organelles known
as bacterial microcompartments (MCPs) (1–8). The function of MCPs is to optimize
metabolic pathways having intermediates that are toxic or poorly retained by the cell
envelope (5, 7, 9). They consist of sequentially acting metabolic enzymes (an enzymatic
core) encapsulated within a protein shell that controls the diffusion of enzyme sub-
strates and products while confining toxic or volatile intermediates (10–14). MCPs are
among the largest multiprotein complexes known. They are typically 100 to 150 nm in
diameter and up to a gigadalton or more in mass and are built from thousands of
protein subunits of 10 to 20 different types. A key question of MCP architecture, and the
subject of this report, is the protein-protein interactions that guide the assembly of
Received 9 November 2016 Accepted 20
January 2017
Accepted manuscript posted online 30
January 2017
Citation Lehman BP, Chowdhury C, Bobik TA.
2017. The N terminus of the PduB protein
binds the protein shell of the Pdu
microcompartment to its enzymatic core. J
Bacteriol 199:e00785-16. https://doi.org/
10.1128/JB.00785-16.
Editor William W. Metcalf, University of Illinois
at Urbana-Champaign
Copyright © 2017 American Society for
Microbiology. All Rights Reserved.
Address correspondence to Thomas A. Bobik,
bobik@iastate.edu.
RESEARCH ARTICLE
crossm
April 2017 Volume 199 Issue 8 e00785-16 jb.asm.org 1Journal of Bacteriology
thousands of protein subunits into a functioning bacterial organelle with a defined
higher-order structure.
The first bacterial MCP identified was the carboxysome, which is used to enhance
autotrophic CO
2
fixation (15, 16). Subsequent studies showed that MCPs are involved
in the catabolism of 1,2-propanediol (1,2-PD), ethanolamine, choline, glycerol, rham-
nose, fucose, and fucoidan (11, 17–22). Based on bioinformatics analyses, MCPs are
produced by about 20% of bacteria distributed across 11 kingdom-level taxa and have
a number of variant forms with uncertain physiological roles (1, 2, 6, 8). Notably, MCPs
have been linked to bacterial pathogenesis, as well as to heart disease, via their
metabolic role in the gut microbiome, and they are a promising basis for the devel-
opment of engineered nanocompartments (5, 23–29). A remarkable feature of bacterial
MCPs is that the outer shells of diverse metabolic types are built primarily from the
same family of proteins known as bacterial microcompartment (BMC) domain proteins
(26, 30, 31). BMC domain family members have a flat hexagonal quaternary structure
and tile edge to edge into extended protein sheets (30). Many are functionally
diversified, and a typical MCP shell is composed of 4 to 10 different BMC domain
proteins (5, 7). The edge-to-edge interactions that drive shell protein tessellation are
thought to be conserved among varied types of BMC domain proteins, allowing the
formation of complex shells (31, 32). Another key aspect of MCP architecture is the use
of short sequence extensions on the N and C termini of both lumen enzymes and
structural proteins to guide assembly (33–37). Many enzymes that localize to the lumen
of bacterial MCPs have N-terminal extensions that are absent from homologs that are
not associated with MCPs (35, 37), and in a number of cases, these extensions (typically
about 20 amino acids in length) are necessary and sufficient for the encapsulation of
enzymes into MCP shells (25, 33–37). In one case, an N-terminal-targeting sequence was
found to interact with short C-terminal
-helix conserved on several shell proteins, and
this binding interaction is thought to guide encapsulation (38, 39). In addition, a short
C-terminal sequence of the CcmN protein is essential for carboxysome formation and
is proposed to bind the CcmK2 shell protein (37).
Perhaps the best understood MCP is the 1,2-PD utilization (Pdu) MCP of Salmonella
enterica serovar Typhimurium LT2. The Pdu MCP encapsulates enzymes used for the
B
12
-dependent degradation of 1,2-PD, which is an important carbon and energy source,
particularly in anaerobic environments (40, 41). Initially, 1,2-PD diffuses into the lumen
of the MCP, where it is converted to propionaldehyde by coenzyme B
12
-dependent diol
dehydratase (40, 42). Propionaldehyde is then converted to propionyl coenzyme A
(propionyl-CoA) by CoA-dependent aldehyde dehydrogenase (PduP) or to 1-propanol
by the PduQ alcohol dehydrogenase (43, 44). The PduL enzyme converts propionyl-CoA
to propionyl-PO
4
, which exits the MCP and enters central metabolism via the methyl
citrate pathway (45–48). The protein shell of the Pdu MCP confines propionaldehyde,
minimizing its toxicity and diffusive loss (14, 49). This shell is composed of seven BMC
domain proteins (PduA, PduB, PduB=, PduJ, PduK, PduT, and PduU) and a bacterial
microcompartment vertex (BMV) protein (PduN) and may also include PduM, but the
function of this protein is uncertain (50–52). The most abundant shell proteins of the
Pdu MCP are PduB and PduB=(51). Together, they make up about 50% of the shell and
almost 25% of the total MCP protein (51). PduB and PduB=are expressed from
overlapping genes and are identical in sequence, except that PduB has a 37-amino-acid
N-terminal extension (51, 53). In this report, we investigate the role of the 37-amino-
acid N-terminal extension of PduB, and based on our results, we present a new model
for MCP assembly.
RESULTS
Multiple-sequence analysis and secondary structure prediction for the
N-terminal region of PduB. To investigate the function of the N-terminal region of
PduB, we constructed a multiple-sequence alignment to look for conserved sequence
features. PSI-BLAST (54) was used to identify PduB homologs in GenBank, and Clustal
Omega (55) was used to construct the alignment. Amino acid residues L6 to V18 of the
Lehman et al. Journal of Bacteriology
April 2017 Volume 199 Issue 8 e00785-16 jb.asm.org 2
PduB N-terminal region were highly conserved across diverse genera, but residues A18
to E35 were much more divergent, particularly when distantly related genera were
compared (Fig. 1). Analyses with the small-peptide structure prediction software PEP-
FOLD (56) suggested that the N terminus of PduB is composed of an
-helix (L6 to V18),
followed by a disordered region (A18 to E35). Thus, the N-terminal extension that
distinguishes PduB from PduB=consists of a highly conserved predicted
-helix and a
less conserved disordered region connected to the two tandem BMC domains that
comprise the core structure of both PduB and PduB=(57).
N-terminal deletion mutants. To further investigate the role of the N-terminal
extension of PduB, a series of scarless mutants were constructed by Sac-Cat recom-
bineering and verified by DNA sequencing (Fig. 2): in the PduB Δ1–37 mutant, the
entire N-terminal extension that distinguishes PduB from PduB=is deleted; in the PduB
Δ6 –12 mutant, the N-terminal half of the highly conserved predicted
-helix is re-
moved; in the PduB Δ11–25 mutant, the C-terminal half of the same
-helix and part
of the linker region are removed; in the PduB Δ27–32 mutant, a somewhat unusual
proline-rich area of the linker region is deleted; and in the PduB M38A mutant, the start
codon for PduB=is changed to CGC (alanine) such that PduB M38A is produced but
PduB=is not (verified as described below).
Effect of N-terminal PduB deletions on MCP assembly. Our initial analysis of the
PduB deletions was to test their effects on MCP assembly in vivo (Fig. 2). Prior studies
showed that Salmonella mutants unable to correctly assemble the Pdu MCP grow
substantially faster than the wild type on 1,2-PD minimal medium with limiting B
12
(49).
This phenotype, which is due to increased access of the MCP lumen enzymes to their
substrates, is a reliable test of the integrity/permeability of the Pdu MCP (49, 50, 58). The
deletion mutants described above showed a range of growth rates on 1,2-PD at limiting
B
12
(Fig. 2). The PduB Δ1–37, PduB Δ6 –12, and PduB Δ11–25 mutants grew much faster
than the wild type on minimal 1,2-PD medium, with limiting B
12
indicating an assembly
defect. Indeed, the growth rates resulting from these short deletions were similar to
that observed for a full-length pduBB=deletion, which was previously shown to produce
no MCPs (50). On the other hand, the PduB Δ27–32 mutation and the PduB M38A
mutation (which prevents translation of the PduB=protein by replacement of its start
codon) resulted in intermediate growth phenotypes indicative of a partial assembly
defect (Fig. 2). We consider it unlikely that any of the mutations described above had
substantial effects on PduB folding, since prior crystallographic studies showed that
PduB lacking 37 N-terminal amino acids (PduB=) folds normally (57) and, below, we
show that these mutant proteins are still incorporated into the shell of the Pdu MCP.
Hence, the studies described above indicate that the N-terminal region of PduB plays
a role in assembly of the Pdu MCP.
FIG 1 Multiple-sequence alignment of the first 53 amino acids of the PduB/PduB=coding region, with their respective
translation start sites indicated. The PduB protein has a 37-amino-acid N-terminal extension not present in PduB=but is
otherwise identical in amino acid sequence. Blue-highlighted residues have 80% amino acid sequence identity across the
sequences shown, whose GI numbers are (top to bottom) 409994423, 545166260, 737633100, 490278627, 507082779,
506359047, 738157884, 696233319, 746121973, and 496088740.
PduB Is Essential for Assembly of Pdu MCP Journal of Bacteriology
April 2017 Volume 199 Issue 8 e00785-16 jb.asm.org 3
To further investigate the role of the N terminus of PduB in MCP assembly, MCPs
from each of the mutants described above were purified and analyzed. MCP yield (the
total protein content of the purified MCPs/gram of cells) relative to the wild type was
used to evaluate MCP stability during purification. SDS-PAGE was used to examine
protein content, and enzyme assays were used to determine the relative abundance of
two major lumen enzymes, B
12
-dependent diol dehydratase (PduCDE) and aldehyde
dehydrogenase (PduP). The PduB M38A and PduB Δ27–32 mutant MCPs were purified
with somewhat reduced yields compared to the wild type, at 66% and 75%, respec-
tively, consistent with the partially impaired assembly/stability indicated by growth
tests (Fig. 3). However, the protein content of these mutant MCPs was similar to the
wild type by SDS-PAGE, with the exception of the changes due to the mutations (Fig.
3). For the PduB Δ27–32 mutant, the truncated PduB protein ran just above the PduD
band. For the PduB M38A mutant, the PduB=band is missing, consistent with prior
studies that indicated that M38 is the start of PduB=(51, 53). All other major proteins
in MCPs purified from these mutants were present at levels similar to the wild type.
Some minor MCP proteins (PduN, PduK, and PduL) are difficult to observe by SDS-
PAGE; hence, their abundance in the mutant MCPs is uncertain. We also measured the
PduCDE (diol dehydratase) activity of the purified mutant MCPs. The activities of
the PduB M38A (25.5 0.9
mol · min
1
·mg
1
) and PduB Δ27–32 (25.6 0.9
mol ·
min
1
·mg
1
) mutants were similar to that of the wild type (26.0 1.3
mol · min
1
·
mg
1
), indicating normal encapsulation of PduCDE within the MCP (Fig. 3). For PduP
aldehyde dehydrogenase, activities were 9.3 0.6, 10.2 0.8, and 10.3 0.08
mol ·
min
1
·mg
1
for the wild type, PduB Δ27–32 mutant, and PduB M38A mutant,
respectively (Fig. 3). Overall, analyses of purified MCPs suggest that the PduB Δ27–32
and PduB M38A mutants have a defect that partially impairs MCP assembly or stability,
leading to a reduced yield during purification; however, this defect does not have a
large effect on the composition of purified MCPs.
FIG 2 Short N-terminal deletions in PduB result in faster growth of Salmonella on 1,2-propanediol
minimal medium. The top panel is a schematic of N-terminal deletions that were tested. Residues
highlighted in red indicate the deleted regions. Underlined residues correspond to a highly conserved
putative
-helix. The pduBB=deletion mutant is a control known to have a very high growth rate due to
an MCP assembly defect. Other PduB deletion mutants are indicated in the legend. The M38A mutant
changes the start codon of PduB=and blocks its translation; thus, it is essentially a ΔpduB=mutant. The
table to the right shows the doubling time for each strain, and the error shown is one standard deviation
determined from the results from three or more independent experiments. The wild type (WT) is
Salmonella enterica LT2 serovar Typhimurium. OD
600
, optical density at 600 nm.
Lehman et al. Journal of Bacteriology
April 2017 Volume 199 Issue 8 e00785-16 jb.asm.org 4
The other N-terminal deletion mutants analyzed (PduB Δ1–37, PduB Δ11–25, and
PduB Δ6 –12) all had growth phenotypes indicative of a more severe MCP assembly
defect (Fig. 2). In addition, yields of MCPs from these mutants (total protein obtained
in the MCP fraction/gram of cells) were drastically lower than that from the wild type,
as only 11%, 27%, and 30% relative recovery was obtained, respectively, which is
indicative of a substantial assembly defect (Fig. 4). We also analyzed the protein content
of the MCP fractions purified from the PduB Δ1–37, PduB Δ11–25, and PduB Δ6 –12
mutants by SDS-PAGE (Fig. 4). Prior studies have shown that the MCP purification
protocol used here allows the isolation of partially assembled and misassembled MCP
complexes (32, 50). The complexes obtained from the PduB Δ1–37, PduB Δ11–25, and
PduB Δ6 –12 mutants had a substantially altered protein composition compared to
wild-type MCPs (Fig. 4). These complexes included the major shell proteins of the Pdu
MCP, PduA, PduB=, and PduJ, as well as truncated versions of PduB resulting from the
mutagenesis and the minor shell proteins PduM, PduU, and PduT. Interestingly, how-
ever, these mutants appeared to have substantially reduced levels of the major lumen
enzymes (PduG, PduC, PduP, PduQ plus PduO, PduD, and PduE) (Fig. 4). Enzyme assays
substantiated that the MCP complexes isolated from the PduB Δ1–37, PduB 6 –12, and
PduB Δ11–25 mutants contained much lower levels of the PduCDE diol dehydratase
and the PduP aldehyde dehydrogenase than the wild type (Fig. 4). We propose that
these mutations impaired the association of lumen enzymes with the MCP shell but had
lesser effects on interactions among the major shell proteins, although more complex
explanations are possible. It is notable that the mutant forms of PduB analyzed here
were incorporated into the MCP shell complexes that were isolated, supporting their
normal folding. Furthermore, the yields of purified “empty shells” were underestimated,
since the enzymatic core constitutes about 50% of the total MCP protein (Fig. 4). Hence,
FIG 3 SDS-PAGE of MCPs purified from the PduB Δ27–32 mutant, the PduB M38A mutant, and the wild
type (WT). Letters on the right indicate the Pdu protein represented by each band. Lysozyme (lys) is not
an MCP component but was used to lyse cells. The percent yield of purified MCPs indicates grams of
protein/gram of cells relative to the wild type. The specific activities of the major lumen enzymes, the
PduCDE diol dehydratase and the PduP aldehyde dehydrogenase, are shown in micromoles per minute
per milligram. Ten micrograms of protein was loaded in each lane. Gels were stained using a Coomassie-
based protocol. The enzyme assay methods are described in Materials and Methods. The error shown is
one standard deviation determined from the results from three or more independent experiments. M,
molecular mass markers.
PduB Is Essential for Assembly of Pdu MCP Journal of Bacteriology
April 2017 Volume 199 Issue 8 e00785-16 jb.asm.org 5
analysis of the MCPs purified from the PduB Δ1–37, PduB Δ11–25, and PduB Δ6 –12
mutants substantiates a key role for the N terminus of PduB in MCP assembly.
Point mutations in the N-terminal
-helix of PduB. To further define the role of
the N terminus of PduB in MCP assembly, the highly conserved predicted
-helical
region (residues L6 to V18) was examined. Thirteen scarless chromosomal mutations
were constructed (L6T, V7T, E8A, Q9A, I10T, M11S, A12S, Q13A, V14T, I15T, R17A, and
V18T), such that each amino acid side chain along the helix was changed from polar to
nonpolar or vice versa. All mutations were verified by DNA sequencing. The effect of
each mutation on MCP integrity was first examined using growth tests. As described
above, fast growth of Salmonella on 1,2-PD at low B
12
indicates a compromised MCP
structure (49, 50). Strains having PduB L6T, A12S, and I15T mutations grew similarly to
the wild type on 1,2-PD minimal medium with low B
12
. This suggested that these
residues do not play a critical role in MCP assembly. Strains with PduB E8A, Q9A, Q13A,
R17A, and V18T mutations grew slightly slower than the wild type on 1,2-PD medium,
but the effect was small and not clearly significant. More interesting were strains with
PduB V7T, I10T, M11S, and V14T mutations, which showed significantly faster growth
than the wild type (Fig. 5). This suggests that V7, I10, M11, and V14 of PduB have
important functional roles in assembly of the Pdu MCP.
To further investigate the effects of the PduB V7T, M11S, V14T, and I10T mutations
on MCP assembly, MCPs were purified and analyzed from each mutant strain. Purifi-
cation yields were 80%, 90%, 70%, and 25% relative to the wild type, respectively,
indicating that each mutation reduced MCP stability, with I10T having the largest effect.
When we examined the protein content of MCPs purified from each mutant by
SDS-PAGE, the expected shell proteins (PduB, PduB=, PduA, PduJ, PduT, PduU, and
PduM) were present in each case (Fig. 6). Notably, however, MCPs purified from the
PduB I10T mutant had substantially reduced levels of lumen enzymes compared to
the wild type, including PduCDE diol dehydratase and PduP aldehyde dehydrogenase
(PduP), and the other mutant MCPs appeared to have somewhat reduced PduCDE
levels (Fig. 6). To better quantify the lumen enzymes present, we measured the PduCDE
and PduP enzyme activities in MCPs purified from each mutant (Fig. 6). PduCDE diol
FIG 4 SDS-PAGE of MCPs purified from pduB deletion mutants with MCP assembly defects. WT, wild type.
Letters on the right indicate the Pdu protein represented by each band. Lysozyme (lys) is not an MCP
component but was used to lyse cells. The percent yield of purified MCPs indicates grams of protein/
gram of cells relative to the wild type. The specific activities of the major lumen enzymes, the PduCDE
diol dehydratase and the PduP aldehyde dehydrogenase, are shown in micromoles per minute per
milligram. The error shown is one standard deviation determined from the results from three or more
independent experiments. Ten micrograms of protein was loaded in each lane. Gels were stained using
a Coomassie-based protocol. Enzyme assays are described in Materials and Methods.
Lehman et al. Journal of Bacteriology
April 2017 Volume 199 Issue 8 e00785-16 jb.asm.org 6
dehydratase activity was lower than the wild type in each mutant MCP, with the I10T
mutant showing the biggest differential (18% activity compared to the wild type)
(Fig. 6). The specific activity of PduP (the second most abundant lumen enzyme) was
reduced in the I10T mutant MCPs (35% activity) but not in the other mutant MCPs
assayed. Thus, analyses of purified MCPs supported the growth studies that indicated
a role for PduB V7, I10, M11, and V14 in proper MCP assembly. They also tentatively
suggested that these residues have particular importance in the association of the
PduCDE diol dehydratase with the Pdu MCP.
FIG 5 Individual point mutations along the highly conserved putative
-helix on the N terminus of PduB
result in fast growth on 1,2-propanediol minimal medium. WT, wild type. The PduB Δ6 –12 mutant is a
control strain known to have a very high growth rate due to an MCP assembly defect. Four pduB mutants
with site-directed mutations (V7T, I19T, M11S, and V14T) were also tested. The table to the right shows
the doubling time for each strain, and the error shown is one standard deviation determined from the
results from three or more independent experiments. Growth curves were determined with a microplate
reader, as described in Materials and Methods.
FIG 6 SDS-PAGE of MCPs purified from PduB mutants with site-directed mutations in the conserved
N-terminal
-helix. WT, wild type. Letters on the right indicate the Pdu protein represented by each band.
Lysozyme (lys) is not an MCP component but was used to lyse cells. The percent yield of purified MCPs
indicates grams of protein/gram of cells relative to the wild type. The specific activities of the major
lumen enzymes, the PduCDE diol dehydratase and the PduP aldehyde dehydrogenase, are given in
micromoles per minute per milligram. The error shown is one standard deviation determined from the
results from three or more independent experiments. Ten micrograms of protein was loaded in each
lane. Gels were stained using a Coomassie-based protocol. Enzyme assays are described in Materials and
Methods. On the right is a model that maps the residues of the PduB N terminus that were altered.
PduB Is Essential for Assembly of Pdu MCP Journal of Bacteriology
April 2017 Volume 199 Issue 8 e00785-16 jb.asm.org 7
Transmission electron microscopy of purified MCPs from selected mutants.
Transmission electron microscopy (TEM) was used to evaluate the MCPs/partially
assembled MCPs purified from selected mutants, including the PduB M38A, PduB
Δ1–37, and PduB Δ6 –12 mutants (Fig. 7). TEM showed that the PduB M38A mutant
formed MCPs similar in appearance to the wild type, which supported the SDS-PAGE
analyses and enzyme assays that indicated a normal protein content (Fig. 3). On the
other hand, TEM of the MCP fraction purified from the PduB Δ1–37 mutant revealed
primarily structures larger than wild-type MCPs, with an appearance similar to empty
MCP shells described earlier (33). This was supportive of the SDS-PAGE results, which
showed that the MCP fraction purified from the PduB Δ1–37 mutant consisted mainly
of shell proteins and only a small amount lumen enzymes (Fig. 4). Similarly, TEM
indicated that the PduB Δ6 –12 mutant formed mainly large (216 66 nm) empty MCP
shells, which was also consistent with SDS-PAGE results (Fig. 4). For both deletion
mutants (PduB Δ1–37 and Δ6 –12), substantial proteinaceous debris was observed in
the purified MCP fraction. Given that SDS-PAGE indicated that these fractions consisted
mainly of shell proteins (Fig. 4), we consider it likely that the debris is shell protein
aggregates. Hence, TEM results support the idea that the N-terminal region of PduB is
needed for proper MCP assembly.
DISCUSSION
Bacterial MCPs are extremely large multiprotein complexes that are widely used as
metabolic organelles. A key question about MCP morphogenesis is the nature of the
FIG 7 Transmission electron microscopy images of MCPs and MCP subcomplexes isolated from selected
pduB deletion mutants. The wild-type strain is Salmonella enterica serovar Typhimurium LT2. MCP/MCP
subcomplexes were negatively stained, as described previously (32, 43). Arrowheads indicate MCPs or
putative empty MCP shells.
Lehman et al. Journal of Bacteriology
April 2017 Volume 199 Issue 8 e00785-16 jb.asm.org 8
interactions that direct the assembly of thousands of protein subunits into a highly
ordered functioning organelle. From a broad perspective, bacterial MCPs consist of a
protein core composed of sequentially acting enzymes encapsulated within a selec-
tively permeable protein shell. Their assembly requires interactions specific to the shell
and the core, as well as associations that tether the shell to the core. In this report, we
present several lines of evidence that indicate that the N-terminal region of PduB plays
a crucial role in binding the shell of the Pdu MCP to its enzymatic core. Various
mutations in the N-terminal region of PduB resulted in a growth phenotype indicative
of impaired MCP assembly (fast growth on 1,2-PD at low B
12
). Purification of MCPs from
several strains carrying mutations in the N terminus of PduB (Δ6 –12, Δ11–25, and
Δ1–37 mutants) led to the isolation of MCPs depleted in the core enzymes, as measured
by SDS-PAGE and enzyme assays, but with the expected content of major shell proteins
as assessed by SDS-PAGE. TEM provided further support, as MCPs purified from two
different mutants revealed structures that appeared to be large empty MCP shells.
Putting this in context, prior studies showed that deletion of any individual shell protein
of the Pdu MCP (other than PduB) did not prevent shell-core interactions, since MCPs
purified from these mutants had the expected complement of core enzymes and shell
proteins (32, 50, 52, 59). Furthermore, we showed that an M38A mutation (which
prevents translation of PduB=) also had relatively little effect on the association of the
shell with the enzymatic core (Fig. 3). Thus, studies encompassing the roles of every Pdu
shell protein in MCP assembly indicate that the N-terminal region of PduB mediates the
major interaction that binds the shell of the Pdu MCP to its enzymatic core.
A theme in the assembly of bacterial MCPs is that short extensions on the N or C
terminus of both structural proteins and core enzymes play a key role in determining
higher-order structure (4). This idea developed from studies that showed the N-terminal
18 amino acids of the PduP core enzyme function as a targeting sequence that is
necessary and sufficient for encapsulation of proteins within the Pdu MCP, as well as an
extensive bioinformatic analysis that predicted that analogous short extensions are
widely used in MCP assembly (35). Further work showed that short N-terminal exten-
sions are used to target enzymes to diverse types of MCPs, both native and recombi-
nant (27, 33, 34, 36, 39, 60–62). Mechanistic studies found that the N-terminal targeting
sequences of PduP form an
-helix that binds to a short C-terminal helix of the PduA,
PduJ, and/or PduK shell proteins to mediate encapsulation (38, 39). Similarly, the
targeting sequences of the EutC and EutE enzymes of the ethanolamine MCP bind a
corresponding helix on the EutS shell protein (although the helix is internal rather than
terminal, due to circular permutation of the EutS shell protein) (62). Thus, it is generally
accepted that binding interactions between enzyme-targeting sequences and shell
proteins drive the encapsulation of enzymes into the lumen of diverse bacterial MCPs.
However, the question of specificity with regard to the binding of the enzymatic core
to the shell is unresolved and somewhat paradoxical. Studies indicate that a single
N-terminal enzyme-targeting sequence can bind multiple shell proteins (38, 39) and
vice versa, i.e., that a single shell protein can bind multiple core enzymes through
targeting sequence associations (59, 62, 63). Recent findings have also shown that
targeting sequences from diverse MCP systems (Eut and Grp) can direct the encapsu-
lation of proteins into the Pdu MCP via a conserved hydrophobic motif (36). Thus, the
seemingly low specificity observed for core-shell interactions raises the question of
whether MCP cores and shells (each of which are composed of 4 to 10 different
proteins) have defined relative molecular orientations. Above, we presented data
indicating that the N-terminal region of PduB plays the primary role in binding the shell
of the Pdu MCP to its enzymatic core. Hence, we propose a model for assembly of the
Pdu MCP in which binding of the N-terminal region of PduB fixes the position of the
shell relative to the core.
In this study, we also conducted a bioinformatic analysis of the N terminus of the
PduB shell protein and found that it consists of a highly conserved putative
-helix and
a less conserved coiled region. Prior studies indicated that the targeting sequences of
core enzymes are short
-helices that bind to short
-helices in shell proteins (38, 62).
PduB Is Essential for Assembly of Pdu MCP Journal of Bacteriology
April 2017 Volume 199 Issue 8 e00785-16 jb.asm.org 9
Binding occurs along conserved surfaces and mediates enzyme encapsulation (36, 38,
60, 62). Site-directed mutagenesis of PduB
-helix at residues 6 to 18 indicated that a
conserved hydrophobic patch formed by residues V7, I10, M11, and V14 was crucial to
MCP assembly. This patch is reminiscent of surfaces present in other MCP assembly
sequences found primarily at the termini of shell proteins and lumen enzymes. Pre-
sumably, this hydrophobic surface interacts with a core enzyme of the Pdu MCP,
possibly a targeting sequence. However, further work will be needed to find the
binding partner(s) of PduB.
A number of labs are developing bacterial MCPs as nanobioreactors for the pro-
duction of renewable chemicals and pharmaceuticals (24, 26, 27, 64). Spatial organiza-
tion of enzymes by encapsulation within protein shells (as exemplified in MCP systems)
has the ability to accelerate catalysis, prevent side reactions, and minimize the harmful
effects of toxic intermediates. A key advance in this field has been the production of
empty MCP shells which were then filled with heterologous cargo by fusing targeting/
encapsulation sequences to desired enzymes/proteins (33, 36, 60–62, 64–68). Further
development of this approach for efficient encapsulation of multiple enzymes with
defined stoichiometries will likely require the use of various types of targeting se-
quences, as well as parameterization of the molecular interactions that define encap-
sulation. Here, we have defined a new terminal sequence (the N-terminal region of the
PduB shell protein) that mediates a major interaction between the enzymatic core and
the shell of the Pdu MCP. This sequence might prove useful for developing additional
systems that mediate protein encapsulation into engineered MCPs.
The findings reported here also relate to two views of MCP biogenesis that have
been proposed. Studies of the
-carboxysome suggest that the shell and the core
assemble more or less simultaneously (69), while studies of the
-carboxysome suggest
that the enzymatic core assembles first and then is rapidly enveloped by the shell
(although the factors that delay association of the shell and core have not been
determined) (70, 71). Given the key role of the N terminus of PduB in fastening the shell
of the Pdu MCP to its enzymatic core, it is likely to play a prominent role in biogenesis,
perhaps even making the first point of contact. We also observed that the essentially
empty MCP shells produced by certain mutants were much larger than wild-type MCPs.
This suggests that interactions between the shell and the core may help to impart
curvature to the shell and define MCP size, but certainly this is not the only factor
involved, and indeed, prior studies have pointed out the key role of vertex proteins in
guiding shell curving (72, 73). Hence, the findings presented here may help with future
studies on the dynamics and mechanisms of MCP assembly.
MATERIALS AND METHODS
Chemicals and reagents. Antibiotics, coenzyme B
12
(Ado-B
12
), vitamin B
12
(CN-B
12
), NADH, and
NAD
were from Sigma Chemical Company (St. Louis, MO). Choice Taq Blue master mix was from
Denville Scientific (Holliston, MA). KOD Hot Start master mix was from EMD Millipore (Billerica, MA).
Bacterial protein extraction reagent (B-PERII), NuPAGE Bis-Tris gels, SimplyBlue SafeStain, and other
reagents were from Thermo Fisher Scientific (Pittsburgh, PA).
Bacterial strains and growth conditions. The bacterial strains used in this study are listed in Table
1. The rich media used were lysogeny broth (LB), also known as Luria-Bertani/Lennox medium (Becton,
Dickinson and Company, Franklin Lakes, NJ) (74), and Terrific broth (MP Biomedicals, Solon, OH). The
minimal medium was no-carbon-E (NCE) medium (75). Growth curves were determined using a Synergy
HT microplate reader (BioTek, Winooski, VT), as previously described (32). NCE minimal medium was
supplemented with 0.6% 1,2-PD, 0.3 mM each Val, Ile, Leu, Thr, Phe, Met, and Cys, 0.1 mM each Tyr and
Trp, 20
M calcium pantothenate, 50
M iron(III) citrate, and CN-B
12
, as indicated in Results.
Construction of chromosomal mutations. Scarless chromosomal deletions and point mutations
were made by Sac-Cat recombineering using the lambda Red recombinase, as described previously (32,
76). The oligonucleotides used for recombineering are listed in Table S1. All mutations were verified by
DNA sequencing.
MCP purification and electron microscopy. MCP purification was performed by detergent extrac-
tion and differential centrifugation, as described previously (52), with the modification that growth
medium was supplemented with 50
M Fe(III) citrate. The amount of purified MCPs obtained (in grams
of protein) per gram of cells was used to calculate the percent yield. For electron microscopy, purified
MCPs were fixed to carbon-coated copper grids and negatively stained, as described previously (32, 43).
Grids were viewed under a JEOL 2100 scanning transmission electron microscope (JEOL USA, Inc.,
Peabody, MA).
Lehman et al. Journal of Bacteriology
April 2017 Volume 199 Issue 8 e00785-16 jb.asm.org 10
Enzyme assays and protein methods. Coenzyme B
12
-dependent diol dehydratase (PduCDE) activ-
ities were determined by a linked continuous spectrophotometric assay, as described previously (59).
PduP aldehyde dehydrogenase activities were measured by following the conversion of NAD
to NADH
spectrophotometrically at 340 nm, as described previously (44). Sodium dodecyl sulfate-polyacrylamide
gel electrophoresis (SDS-PAGE) was carried out using NuPAGE 4 to 12% Bis-Tris gels, Novex mini cells
(Invitrogen), and SimplyBlue SafeStain (Thermo Fisher Scientific). Protein was measured using Bio-Rad
protein assay reagent, which is based on the method of Bradford, with bovine serum albumin as the
standard.
ACKNOWLEDGMENTS
This work was supported by grant AI081146 from the National Institutes of Health
to T.A.B.
We thank the ISU DNA Sequencing and Synthesis Facility for assistance with DNA
analyses, the ISU Microscopy and Nanoimaging Facility for help with electron micros-
copy, and members of the Bobik laboratory, especially Sharmistha Sinha, for helpful
discussions.
REFERENCES
1. Abdul-Rahman F, Petit E, Blanchard JL. 2013. The distribution of poly-
hedral bacterial microcompartments suggests frequent horizontal trans-
fer and operon reassembly. J Phylogen Evol Biol 1:118. https://doi.org/
10.4172/2329-9002.1000118.
2. Axen SD, Erbilgin O, Kerfeld CA. 2014. A taxonomy of bacterial micro-
compartment loci constructed by a novel scoring method. PLoS Comput
Biol 10:e1003898. https://doi.org/10.1371/journal.pcbi.1003898.
3. Bobik TA. 2006. Polyhedral organelles compartmenting bacterial meta-
bolic processes. Appl Microbiol Biotechnol 70:517–525. https://doi.org/
10.1007/s00253-005-0295-0.
4. Bobik TA, Lehman BP, Yeates TO. 2015. Bacterial microcompartments:
widespread prokaryotic organelles for isolation and optimization of
metabolic pathways. Mol Microbiol 98:193–207. https://doi.org/10.1111/
mmi.13117.
5. Chowdhury C, Sinha S, Chun S, Yeates TO, Bobik TA. 2014. Diverse
bacterial microcompartment organelles. Microbiol Mol Biol Rev 78:
438 468. https://doi.org/10.1128/MMBR.00009-14.
6. Jorda J, Lopez D, Wheatley NM, Yeates TO. 2013. Using comparative
genomics to uncover new kinds of protein-based metabolic organelles
in bacteria. Protein Sci 22:179 –195. https://doi.org/10.1002/pro.2196.
7. Rae BD, Long BM, Badger MR, Price GD. 2013. Functions, compositions,
and evolution of the two types of carboxysomes: polyhedral microcom-
partments that facilitate CO
2
fixation in cyanobacteria and some pro-
teobacteria. Microbiol Mol Biol Rev 77:357–379. https://doi.org/10.1128/
MMBR.00061-12.
8. Zarzycki J, Erbilgin O, Kerfeld CA. 2015. Bioinformatic characterization of
glycyl radical enzyme-associated bacterial microcompartments. Appl
Environ Microbiol 81:8315– 8329. https://doi.org/10.1128/AEM.02587-15.
9. Kerfeld CA, Heinhorst S, Cannon GC. 2010. Bacterial microcompartments.
Annu Rev Microbiol 64:391– 408. https://doi.org/10.1146/annurev.micro
.112408.134211.
10. Dou Z, Heinhorst S, Williams EB, Murin CD, Shively JM, Cannon GC. 2008.
CO
2
fixation kinetics of Halothiobacillus neapolitanus mutant carboxy-
somes lacking carbonic anhydrase suggest the shell acts as a diffusional
barrier for CO
2
. J Biol Chem 283:10377–10384. https://doi.org/10.1074/
jbc.M709285200.
11. Penrod JT, Roth JR. 2006. Conserving a volatile metabolite: a role for
carboxysome-like organelles in Salmonella enterica. J Bacteriol 188:
2865–2874. https://doi.org/10.1128/JB.188.8.2865-2874.2006.
12. Rondon MR, Horswill AR, Escalante-Semerena JC. 1995. DNA polymerase
I function is required for the utilization of ethanolamine, 1,2-
TABLE 1 Strains used in this study
Strain
a
Description Source
BE293 Carries pKD46 Lab collection
SS23 ΔpduBB=Lab collection
BL55 ΔpduB (P27–P32)::pblaP3-sacB-cat/pKD46 This study
BL56 ΔpduB (P27–P32) This study
BL62 ΔpduB (M1–A37)::pblaP3-sacB-cat/pKD46 This study
BL137 pduB M38A This study
BL144 ΔpduB (M11–A25) This study
BL155 ΔpduB (L6–A12)::pblaP3-sacB-cat/pkD46 This study
BL156 ΔpduB (L6–A12) This study
BL157 pduB (L6T) This study
BL158 pduB (E8A) This study
BL159 pduB (L10T) This study
BL160 pduB (A12S) This study
BL179 pduB (V7T) This study
BL180 pduB (M11S) This study
BL181 pduB (V14T) This study
BL182 pduB (I15T) This study
BL183 pduB (Q9A) This study
BL184 pduB (Q13A) This study
BL232 ΔpduB (M1–A37) (RBS AGGA for pduB=) This study
BL234 pduB (R17A) This study
BL235 pduB (V18T) This study
a
All strains are derivatives of S. enterica serovar Typhimurium LT2 which contain scarless mutations that
change the amino acid/nucleotide sequence as indicated. RBS, ribosome binding site.
PduB Is Essential for Assembly of Pdu MCP Journal of Bacteriology
April 2017 Volume 199 Issue 8 e00785-16 jb.asm.org 11
propanediol, and propionate by Salmonella Typhimurium LT2. J Bacte-
riol 177:7119 –7124. https://doi.org/10.1128/jb.177.24.7119-7124.1995.
13. Rondon MR, Kazmierczak R, Escalante-Semerena JC. 1995. Glutathi-
one is required for maximal transcription of the cobalamin biosyn-
thetic and 1,2-propanediol utilization (cob/pdu) regulon and for the
catabolism of ethanolamine, 1,2-propanediol, and propionate in Sal-
monella Typhimurium LT2. J Bacteriol 177:5434 –5439. https://
doi.org/10.1128/jb.177.19.5434-5439.1995.
14. Sampson EM, Bobik TA. 2008. Microcompartments for B
12
-dependent
1,2-propanediol degradation provide protection from DNA and cellular
damage by a reactive metabolic intermediate. J Bacteriol 190:
2966 –2971. https://doi.org/10.1128/JB.01925-07.
15. Price GD, Badger MR. 1989. Isolation and characterization of high CO
2
-
requiring-mutants of the cyanobacterium Synechococcus PCC7942: two
phenotypes that accumulate inorganic carbon but are apparently un-
able to generate CO
2
within the carboxysome. Plant Physiol 91:514 –525.
https://doi.org/10.1104/pp.91.2.514.
16. Shively JM, Ball F, Brown DH, Saunders RE. 1973. Functional organelles in
prokaryotes: polyhedral inclusions (carboxysomes) of Thiobacillus nea-
politanus. Science 182:584 –586. https://doi.org/10.1126/science
.182.4112.584.
17. Craciun S, Balskus EP. 2012. Microbial conversion of choline to trimeth-
ylamine requires a glycyl radical enzyme. Proc Natl Acad SciUSA
109:21307–21312. https://doi.org/10.1073/pnas.1215689109.
18. Erbilgin O, McDonald KL, Kerfeld CA. 2014. Characterization of a planc-
tomycetal organelle: a novel bacterial microcompartment for the aero-
bic degradation of plant saccharides. Appl Environ Microbiol 80:
2193–2205. https://doi.org/10.1128/AEM.03887-13.
19. Jameson E, Fu T, Brown IR, Paszkiewicz K, Purdy KJ, Frank S, Chen Y.
2016. Anaerobic choline metabolism in microcompartments promotes
growth and swarming of Proteus mirabilis. Environ Microbiol 18:
2886 –2898. https://doi.org/10.1111/1462-2920.13059.
20. Martínez-del Campo A, Bodea S, Hamer HA, Marks JA, Haiser HJ, Turn-
baugh PJ, Balskus EP. 2015. Characterization and detection of a widely
distributed gene cluster that predicts anaerobic choline utilization by
human gut bacteria. mBio 6(2):e00042–15. https://doi.org/10.1128/
mBio.00042-15.
21. Sriramulu DD, Liang M, Hernandez-Romero D, Raux-Deery E, Lunsdorf H,
Parsons JB, Warren MJ, Prentice MB. 2008. Lactobacillus reuteri DSM
20016 produces cobalamin-dependent diol dehydratase in metabolo-
somes and metabolizes 1,2-propanediol by disproportionation. J Bacte-
riol 190:4559 4567. https://doi.org/10.1128/JB.01535-07.
22. Talarico TL, Casas IA, Chung TC, Dobrogosz WJ. 1988. Production and
isolation of reuterin, a growth inhibitor produced by Lactobacillus reuteri.
Antimicrob Agents Chemother 32:1854 –1858. https://doi.org/10.1128/
AAC.32.12.1854.
23. Conner CP, Heithoff DM, Julio SM, Sinsheimer RL, Mahan MJ. 1998.
Differential patterns of acquired virulence genes distinguish Salmonella
strains. Proc Natl Acad SciUSA95:4641– 4645. https://doi.org/10.1073/
pnas.95.8.4641.
24. Frank S, Lawrence AD, Prentice MB, Warren MJ. 2013. Bacterial micro-
compartments moving into a synthetic biological world. J Biotechnol
163:273–279. https://doi.org/10.1016/j.jbiotec.2012.09.002.
25. Held M, Quin MB, Schmidt-Dannert C. 2013. Eut bacterial
microcompartments: insights into their function, structure, and bioen-
gineering applications. J Mol Microbiol Biotechnol 23:308 –320. https://
doi.org/10.1159/000351343.
26. Kerfeld CA, Erbilgin O. 2015. Bacterial microcompartments and the
modular construction of microbial metabolism. Trends Microbiol 23:
22–34. https://doi.org/10.1016/j.tim.2014.10.003.
27. Kim EY, Tullman-Ercek D. 2013. Engineering nanoscale protein compart-
ments for synthetic organelles. Curr Opin Biotechnol 24:627– 632.
https://doi.org/10.1016/j.copbio.2012.11.012.
28. Thiennimitr P, Winter SE, Winter MG, Xavier MN, Tolstikov V, Huseby DL,
Sterzenbach T, Tsolis RM, Roth JR, Baumler AJ. 2011. Intestinal inflam-
mation allows Salmonella to use ethanolamine to compete with the
microbiota. Proc Natl Acad SciUSA108:17480 –17485. https://doi.org/
10.1073/pnas.1107857108.
29. Tsai SJ, Yeates TO. 2011. Bacterial microcompartments insights into the
structure, mechanism, and engineering applications. Prog Mol Biol Transl
Sci 103:1–20. https://doi.org/10.1016/B978-0-12-415906-8.00008-X.
30. Kerfeld CA, Sawaya MR, Tanaka S, Nguyen CV, Phillips M, Beeby M,
Yeates TO. 2005. Protein structures forming the shell of primitive bac-
terial organelles. Science 309:936 –938. https://doi.org/10.1126/
science.1113397.
31. Yeates TO, Jorda J, Bobik TA. 2013. The shells of BMC-type microcom-
partment organelles in bacteria. J Mol Microbiol Biotechnol 23:290 –299.
https://doi.org/10.1159/000351347.
32. Sinha S, Cheng S, Sung YW, McNamara DE, Sawaya MR, Yeates TO, Bobik
TA. 2014. Alanine scanning mutagenesis identifies an asparagine-
arginine-lysine triad essential to assembly of the shell of the Pdu micro-
compartment. J Mol Biol 426:2328 –2345. https://doi.org/10.1016/
j.jmb.2014.04.012.
33. Choudhary S, Quin MB, Sanders MA, Johnson ET, Schmidt-Dannert C. 2012.
Engineered protein nano-compartments for targeted enzyme localization.
PLoS One 7:e33342. https://doi.org/10.1371/journal.pone.0033342.
34. Fan C, Bobik TA. 2011. The N-terminal region of the medium subunit
(PduD) packages adenosylcobalamin-dependent diol dehydratase
(PduCDE) into the Pdu microcompartment. J Bacteriol 193:5623–5628.
https://doi.org/10.1128/JB.05661-11.
35. Fan C, Cheng S, Liu Y, Escobar CM, Crowley CS, Jefferson RE, Yeates TO,
Bobik TA. 2010. Short N-terminal sequences package proteins into bac-
terial microcompartments. Proc Natl Acad SciUSA107:7509 –7514.
https://doi.org/10.1073/pnas.0913199107.
36. Jakobson CM, Kim EY, Slininger MF, Chien A, Tullman-Ercek D. 2015.
Localization of proteins to the 1,2-propanediol utilization microcompart-
ment by non-native signal sequences is mediated by a common hydro-
phobic motif. J Biol Chem 290:24519 –24533. https://doi.org/10.1074/
jbc.M115.651919.
37. Kinney JN, Salmeen A, Cai F, Kerfeld CA. 2012. Elucidating essential role
of conserved carboxysomal protein CcmN reveals common feature of
bacterial microcompartment assembly. J Biol Chem 287:17729 –17736.
https://doi.org/10.1074/jbc.M112.355305.
38. Fan C, Cheng S, Sinha S, Bobik TA. 2012. Interactions between the
termini of lumen enzymes and shell proteins mediate enzyme encap-
sulation into bacterial microcompartments. Proc Natl Acad SciUSA
109:14995–15000. https://doi.org/10.1073/pnas.1207516109.
39. Lawrence AD, Frank S, Newnham S, Lee MJ, Brown IR, Xue WF, Rowe ML,
Mulvihill DP, Prentice MB, Howard MJ, Warren MJ. 2014. Solution struc-
ture of a bacterial microcompartment targeting peptide and its appli-
cation in the construction of an ethanol bioreactor. ACS Synth Biol
3:454 465. https://doi.org/10.1021/sb4001118.
40. Bobik TA, Havemann GD, Busch RJ, Williams DS, Aldrich HC. 1999. The
propanediol utilization (pdu) operon of Salmonella enterica serovar Ty-
phimurium LT2 includes genes necessary for formation of polyhedral
organelles involved in coenzyme B
12
-dependent 1,2-propanediol deg-
radation. J Bacteriol 181:5967–5975.
41. Obradors N, Badia J, Baldoma L, Aguilar J. 1988. Anaerobic metabolism
of the L-rhamnose fermentation product 1,2-propanediol in Salmonella
Typhimurium. J Bacteriol 170:2159 –2162. https://doi.org/10.1128/
jb.170.5.2159-2162.1988.
42. Bobik TA, Xu Y, Jeter RM, Otto KE, Roth JR. 1997. Propanediol utilization
genes (pdu)ofSalmonella Typhimurium: three genes for the propanediol
dehydratase. J Bacteriol 179:6633– 6639. https://doi.org/10.1128/
jb.179.21.6633-6639.1997.
43. Cheng S, Fan C, Sinha S, Bobik TA. 2012. The PduQ enzyme is an alcohol
dehydrogenase used to recycle NAD
internally within the Pdu micro-
compartment of Salmonella enterica. PLoS One 7:e47144. https://
doi.org/10.1371/journal.pone.0047144.
44. Leal NA, Havemann GD, Bobik TA. 2003. PduP is a coenzyme-A-acylating
propionaldehyde dehydrogenase associated with the polyhedral bodies
involved in B
12
-dependent 1,2-propanediol degradation by Salmonella
enterica serovar Typhimurium LT2. Arch Microbiol 180:353–361. https://
doi.org/10.1007/s00203-003-0601-0.
45. Horswill AR, Escalante-Semerena JC. 1999. Salmonella Typhimurium LT2
catabolizes propionate via the 2-methylcitric acid cycle. J Bacteriol 181:
5615–5623.
46. Liu Y, Jorda J, Yeates TO, Bobik TA. 2015. The PduL phosphotransacylase
is used to recycle coenzyme A within the Pdu microcompartment. J
Bacteriol 197:2392–2399. https://doi.org/10.1128/JB.00056-15.
47. Liu Y, Leal NA, Sampson EM, Johnson CL, Havemann GD, Bobik TA. 2007.
PduL is an evolutionarily distinct phosphotransacylase involved in B
12
-
dependent 1,2-propanediol degradation by Salmonella enterica serovar
Typhimurium LT2. J Bacteriol 189:1589 –1596. https://doi.org/10.1128/
JB.01151-06.
48. Palacios S, Starai VJ, Escalante-Semerena JC. 2003. Propionyl coenzyme
A is a common intermediate in the 1,2-propanediol and propionate
Lehman et al. Journal of Bacteriology
April 2017 Volume 199 Issue 8 e00785-16 jb.asm.org 12
catabolic pathways needed for expression of the prpBCDE operon during
growth of Salmonella enterica on 1,2-propanediol. J Bacteriol 185:
2802–2810. https://doi.org/10.1128/JB.185.9.2802-2810.2003.
49. Havemann GD, Sampson EM, Bobik TA. 2002. PduA is a shell protein
of polyhedral organelles involved in coenzyme B
12
-dependent deg-
radation of 1,2-propanediol in Salmonella enterica serovar Typhimu-
rium LT2. J Bacteriol 184:1253–1261. https://doi.org/10.1128/JB.184
.5.1253-1261.2002.
50. Cheng S, Sinha S, Fan C, Liu Y, Bobik TA. 2011. Genetic analysis of the
protein shell of the microcompartments involved in coenzyme B
12
-
dependent 1,2-propanediol degradation by Salmonella. J Bacteriol 193:
1385–1392. https://doi.org/10.1128/JB.01473-10.
51. Havemann GD, Bobik TA. 2003. Protein content of polyhedral organelles
involved in coenzyme B
12
-dependent degradation of 1,2-propanediol in
Salmonella enterica serovar Typhimurium LT2. J Bacteriol 185:
5086 –5095. https://doi.org/10.1128/JB.185.17.5086-5095.2003.
52. Sinha S, Cheng S, Fan C, Bobik TA. 2012. The PduM protein is a structural
component of the microcompartments involved in coenzyme B
12
-
dependent 1,2-propanediol degradation by Salmonella enterica. J Bac-
teriol 194:1912–1918. https://doi.org/10.1128/JB.06529-11.
53. Parsons JB, Dinesh SD, Deery E, Leech HK, Brindley AA, Heldt D, Frank S,
Smales CM, Lunsdorf H, Rambach A, Gass MH, Bleloch A, McClean KJ,
Munro AW, Rigby SE, Warren MJ, Prentice MB. 2008. Biochemical and
structural insights into bacterial organelle form and biogenesis. J Biol
Chem 283:14366 –14375. https://doi.org/10.1074/jbc.M709214200.
54. Altschul SF, Madden TL, Schaffer AA, Zhang J, Zhang Z, Miller W, Lipman
DJ. 1997. Gapped BLAST and PSI-BLAST: a new generation of protein
database search programs. Nucleic Acids Res 25:3389 –3402. https://
doi.org/10.1093/nar/25.17.3389.
55. McWilliam H, Li W, Uludag M, Squizzato S, Park YM, Buso N, Cowley AP,
Lopez R. 2013. Analysis tool Web Services from the EMBL-EBI. Nucleic
Acids Res 41:W597–W600. https://doi.org/10.1093/nar/gkt376.
56. Maupetit J, Derreumaux P, Tuffery P. 2009. PEP-FOLD: an online resource
for de novo peptide structure prediction. Nucleic Acids Res 37:
W498 –W503. https://doi.org/10.1093/nar/gkp323.
57. Pang A, Liang M, Prentice MB, Pickersgill RW. 2012. Substrate channels
revealed in the trimeric Lactobacillus reuteri bacterial microcompartment
shell protein PduB. Acta Crystallogr D Biol Crystallogr 68:1642–1652.
https://doi.org/10.1107/S0907444912039315.
58. Chowdhury C, Chun S, Pang A, Sawaya MR, Sinha S, Yeates TO, Bobik TA.
2015. Selective molecular transport through the protein shell of a bac-
terial microcompartment organelle. Proc Natl Acad SciUSA112:
2990 –2995. https://doi.org/10.1073/pnas.1423672112.
59. Chowdhury C, Chun S, Sawaya MR, Yeates TO, Bobik TA. 2016. The
function of the PduJ microcompartment shell protein is determined by
the genomic position of its encoding gene. Mol Microbiol 101:770 –783.
https://doi.org/10.1111/mmi.13423.
60. Kim EY, Tullman-Ercek D. 2014. A rapid flow cytometry assay for the
relative quantification of protein encapsulation into bacterial micro-
compartments. Biotechnol J 9:348 –354. https://doi.org/10.1002/
biot.201300391.
61. Lee MJ, Brown IR, Juodeikis R, Frank S, Warren MJ. 2016. Employing
bacterial microcompartment technology to engineer a shell-free
enzyme-aggregate for enhanced 1,2-propanediol production in Esche-
richia coli. Metab Eng 36:48 –56. https://doi.org/10.1016/
j.ymben.2016.02.007.
62. Quin MB, Perdue SA, Hsu SY, Schmidt-Dannert C. 2016. Encapsulation of
multiple cargo proteins within recombinant Eut nanocompartments.
Appl Microbiol Biotechnol 100:9187–9200. https://doi.org/10.1007/
s00253-016-7737-8.
63. Jorda J, Liu Y, Bobik TA, Yeates TO. 2015. Exploring bacterial organelle
interactomes: a model of the protein-protein interaction network in the
Pdu microcompartment. PLoS Comput Biol 11:e1004067. https://
doi.org/10.1371/journal.pcbi.1004067.
64. Bonacci W, Teng PK, Afonso B, Niederholtmeyer H, Grob P, Silver PA,
Savage DF. 2012. Modularity of a carbon-fixing protein organelle.
Proc Natl Acad SciUSA109:478 483. https://doi.org/10.1073/pnas
.1108557109.
65. Held M, Kolb A, Perdue S, Hsu SY, Bloch SE, Quin MB, Schmidt-Dannert
C. 2016. Engineering formation of multiple recombinant Eut protein
nanocompartments in E. coli. Sci Rep 6:24359. https://doi.org/10.1038/
srep24359.
66. Jakobson CM, Chen Y, Slininger MF, Valdivia E, Kim EY, Tullman-Ercek D.
2016. Tuning the catalytic activity of subcellular nanoreactors. J Mol Biol
428:2989 –2996. https://doi.org/10.1016/j.jmb.2016.07.006.
67. Menon BB, Dou Z, Heinhorst S, Shively JM, Cannon GC. 2008. Halothio-
bacillus neapolitanus carboxysomes sequester heterologous and chime-
ric RubisCO species. PLoS One 3:e3570. https://doi.org/10.1371/
journal.pone.0003570.
68. Parsons JB, Frank S, Bhella D, Liang M, Prentice MB, Mulvihill DP, Warren
MJ. 2010. Synthesis of empty bacterial microcompartments, directed
organelle protein incorporation, and evidence of filament-associated
organelle movement. Mol Cell 38:305–315. https://doi.org/10.1016/
j.molcel.2010.04.008.
69. Iancu CV, Morris DM, Dou Z, Heinhorst S, Cannon GC, Jensen GJ. 2010.
Organization, structure, and assembly of alpha-carboxysomes deter-
mined by electron cryotomography of intact cells. J Mol Biol 396:
105–117. https://doi.org/10.1016/j.jmb.2009.11.019.
70. Cameron JC, Wilson SC, Bernstein SL, Kerfeld CA. 2013. Biogenesis of a
bacterial organelle: the carboxysome assembly pathway. Cell 155:
1131–1140. https://doi.org/10.1016/j.cell.2013.10.044.
71. Chen AH, Robinson-Mosher A, Savage DF, Silver PA, Polka JK. 2013. The
bacterial carbon-fixing organelle is formed by shell envelopment of
preassembled cargo. PLoS One 8:e76127. https://doi.org/10.1371/
journal.pone.0076127.
72. Tanaka S, Kerfeld CA, Sawaya MR, Cai F, Heinhorst S, Cannon GC, Yeates
TO. 2008. Atomic-level models of the bacterial carboxysome shell. Sci-
ence 319:1083–1086. https://doi.org/10.1126/science.1151458.
73. Wheatley NM, Gidaniyan SD, Liu Y, Cascio D, Yeates TO. 2013. Bacterial
microcompartment shells of diverse functional types possess pentameric
vertex proteins. Protein Sci 22:660 –665. https://doi.org/10.1002/pro.2246.
74. Bertani G. 1951. Studies on lysogenesis. I. The mode of phage liberation
by lysogenic Escherichia coli. J Bacteriol 62:293–300.
75. Berkowitz D, Hushon JM, Whitfield HJ, Jr, Roth J, Ames BN. 1968.
Procedure for identifying nonsense mutations. J Bacteriol 96:215–220.
76. Datsenko KA, Wanner BL. 2000. One-step inactivation of chromosomal
genes in Escherichia coli K-12 using PCR products. Proc Natl Acad Sci
U S A 97:6640 –6645. https://doi.org/10.1073/pnas.120163297.
PduB Is Essential for Assembly of Pdu MCP Journal of Bacteriology
April 2017 Volume 199 Issue 8 e00785-16 jb.asm.org 13
... Furthermore, the EP of PduP has also been suggested to interact with another BMC-H protein, PduK (37). In separate studies, the N terminus of PduB has been speculated to act as an anchor between the inner cargo and the shell (27,38). If true, we reasoned that purified EPs may associate directly with some or all shell protein superstructures, which could be observed as colocalization events. ...
... From this, we speculate that the N-terminal portion of PduB may facilitate inwardfacing cargo interactions. This is consistent with recent cross-linking mass spectrometry experimentation, which provided direct evidence for interactions between PduB and numerous EP-containing cargo proteins in a native context, as well as other data suggesting that PduB is a shell-cargo anchor point (27,38). PduA nanotubes were similarly found to colocalize with PduD EP and PduP EP (Fig. 6A). ...
... Together, these data tentatively support interactions between EPs and the N terminus of PduB and between EPs and PduA. These data are consistent with the hypothesis that PduB (38,39) and PduA are likely to be primary factors in facilitating shell-cargo interactions. ...
Article
Full-text available
The shell proteins that comprise bacterial microcompartments (BMCs) can self-assemble into an array of superstructures such as nanotubes, flat sheets, and icosahedra. The physical characterization of BMCs and these superstructures typically relies on electron microscopy, which decouples samples from their solution context. We hypothesize that an investigation of fluorescently tagged BMCs and shell protein superstructures in vitro using high-resolution confocal microscopy will lead to new insights into the solution behavior of these entities. We find that confocal imaging is able to capture nanotubes and sheets previously reported by transmission electron microscopy (TEM). Using a combination of fluorescent tags, we present qualitative evidence that these structures intermix with one another in a hetero- and homotypic fashion. Complete BMCs are also able to accomplish intermixing as evidenced by colocalization data. Finally, a simple colocalization experiment suggests that fluorescently modified encapsulation peptides (EPs) may prefer certain shell protein binding partners. Together, these data demonstrate that high-resolution confocal microscopy is a powerful tool for investigating microcompartment-related structures in vitro, particularly for colocalization analyses. These results also support the notion that BMCs may intermix protein components, presumably from the outer shell. IMPORTANCE Microcompartments are large, organelle-like structures that help bacteria catabolize targeted metabolites while also protecting the cytosol against highly reactive metabolic intermediates. Their protein shell self-assembles into a polyhedral structure of approximately 100 to 200 nm in diameter. Inside the shell are thousands of copies of cargo enzymes, which are responsible for a specific metabolic pathway. While different approaches have revealed high-resolution structures of individual microcompartment proteins, it is less clear how these factors self-assemble to form the full native structure. In this study, we show that laser scanning confocal microscopy can be used to study microcompartment proteins. We find that this approach allows researchers to investigate the interactions and potential exchange of shell protein subunits in solution. From this, we conclude that confocal microscopy offers advantages for studying the in vitro structures of other microcompartments as well as carboxysomes and other bacterial organelles.
... Incorporating simulations to complement emerging experimental insights will lead to more meaningful outputs to inform design choices. For instance, future modeling may wish to explore evidenceinformed shell-cargo interaction sites that form from predominantly (i) the edge-edge interaction surface of two adjoining shell proteins (Ni et al., 2023) and, in some cases, (ii) interactions with specific interiororiented domains such as the N-terminus of the PduB BMC-T (Trettel et al., 2022;Lehman et al., 2017;Kennedy et al., 2022). This is further underpinned by the multitude of different shell proteins BMCs can encode and their synthetic interchangeability (Cai et al., 2015b;Slininger Lee et al., 2017) which certainly influence shell-shell (including curvature) and shell-cargo/scaffold interactions. ...
Article
Full-text available
The carboxysome is a bacterial microcompartment (BMC) which plays a central role in the cyanobacterial CO2-concentrating mechanism. These proteinaceous structures consist of an outer protein shell that partitions Rubisco and carbonic anhydrase from the rest of the cytosol, thereby providing a favorable microenvironment that enhances carbon fixation. The modular nature of carboxysomal architectures makes them attractive for a variety of biotechnological applications such as carbon capture and utilization. In silico approaches, such as molecular dynamics (MD) simulations, can support future carboxysome redesign efforts by providing new spatio-temporal insights on their structure and function beyond in vivo experimental limitations. However, specific computational studies on carboxysomes are limited. Fortunately, all BMC (including the carboxysome) are highly structurally conserved which allows for practical inferences to be made between classes. Here, we review simulations on BMC architectures which shed light on (1) permeation events through the shell and (2) assembly pathways. These models predict the biophysical properties surrounding the central pore in BMC-H shell subunits, which in turn dictate the efficiency of substrate diffusion. Meanwhile, simulations on BMC assembly demonstrate that assembly pathway is largely dictated kinetically by cargo interactions while final morphology is dependent on shell factors. Overall, these findings are contextualized within the wider experimental BMC literature and framed within the opportunities for carboxysome redesign for biomanufacturing and enhanced carbon fixation.
... In order to improve the efficiency of intracellular catalytic synthesis, we focused on the formation mechanism of the bacterial intracellular multienzyme complexes. Research on the formation of a bacterial intracellular multi-enzyme complex system has mainly focused on the multi-enzyme complex in the bacterial microcompartment and artificial selfassembled proteins [27][28][29]. Rae and Kerfeld et al. found two types of capsid proteins, α and β, that could form a spherical carboxylase body in Cyanobacteria. With the help of the carboxylase body, different assembly methods were used to combine carbonic anhydrase and ribulose-1,5-bisphosphate carboxylase/oxygenase, which was encapsulated in it and assembled into a multi-enzyme complex [30,31]. ...
Article
Full-text available
Cellulosome is a highly efficient multi-enzyme self-assembly system and is found on the extracellular surface or in the free environment of microorganisms. However, with a lack of Ca2+ in vivo, cellulosome assembly is challenging. In this study, a novel design method was used to directionally modify the Ca2+-binding site, and four double-site dockerin A (DocA) mutants were obtained. At a Ca2+ concentration between 1.00 × 10−7 and 1.00 × 10−4 M, the mutant DocA-D3 had the strongest binding capacity to cohesion (Coh), which was 8.01 times that of DocA. The fluorescence signal intensity of the fusion proteins assembled using mutants was up to 1.26 × 107 in Escherichia coli, which indicated that these mutants could interact with Coh in vivo. The molecular dynamics simulation results showed that DocA-D3 could maintain a stable angle structure without Ca2+, and when applied to L-lysine fermentation, the yield was increased by 24.1%; when applied to β-alanine fermentation, the product accumulation was increased by 2.13–2.63 times. These findings lay the foundation for assembly design in cells.
Article
Bacterial microcompartments (MCPs) are proteinaceous organelles that natively encapsulate the enzymes, substrates, and cofactors within a protein shell. They optimize the reaction rates by enriching the substrate in the vicinity of enzymes to increase the yields of the product and mitigate the outward diffusion of the toxic or volatile intermediates. The shell protein subunits of MCP shell are selectively permeable and have specialized pores for the selective inward diffusion of substrates and products release. Given their attributes, MCPs have been recently explored as potential candidates as subcellular nano-bioreactor for the enhanced production of industrially important molecules by exercising pathway encapsulation. In the current study, MCPs have been shown to sustain enzyme activity for extended periods, emphasizing their durability against a range of physical challenges such as temperature, pH and organic solvents. The significance of an intact shell in conferring maximum protection is highlighted by analyzing the differences in enzyme activities inside the intact and broken shell. Moreover, a minimal synthetic shell was designed with recruitment of a heterologous enzyme cargo to demonstrate the improved durability of the enzyme. The encapsulated enzyme was shown to be more stable than its free counterpart under the aforementioned conditions. Bacterial MCP-mediated encapsulation can serve as a potential strategy to shield the enzymes used under extreme conditions by maintaining the internal microenvironment and enhancing their cycle life, thereby opening new means for stabilizing, and reutilizing the enzymes in several bioprocess industries.
Article
An important goal of systems and synthetic biology is to produce high value chemical species in large quantities. Microcompartments, which are protein nanoshells encapsulating catalytic enzyme cargo, could potentially function as tunable nanobioreactors inside and outside cells to generate these high value species. Modifying the morphology of microcompartments through genetic engineering of shell proteins is one viable strategy to tune cofactor and metabolite access to encapsulated enzymes. However, this is a difficult task without understanding how changing interactions between the many different types of shell proteins and enzymes affect microcompartment assembly and shape. Here, we use multiscale molecular dynamics and experimental data to describe assembly pathways available to microcompartments composed of multiple types of shell proteins with varied interactions. As the average interaction between the enzyme cargo and the multiple types of shell proteins is weakened, the shell assembly pathway transitions from (i) nucleating on the enzyme cargo to (ii) nucleating in the bulk and then binding the cargo as it grows to (iii) an empty shell. Atomistic simulations and experiments using the 1,2-propanediol utilization microcompartment system demonstrate that shell protein interactions are highly varied and consistent with our multicomponent, coarse-grained model. Furthermore, our results suggest that intrinsic bending angles control the size of these microcompartments. Overall, our simulations and experiments provide guidance to control microcomparmtent size and assembly by modulating the interactions between shell proteins.
Article
Full-text available
The self-assembly of bacterial microcompartments is the result of several genetic, biochemical, and physical stimuli orchestrating inside the bacterial cell. In this work, we use 1,2-propanediol utilization microcompartments as a paradigm to identify the factors that physically drive the self-assembly of MCP proteins in vitro using its major shell protein and major encapsulated enzyme. We find that a major shell protein PduBB’ tend to self-assemble under macromolecular crowded environment and suitable ionic strength. Microscopic visualization and biophysical studies reveal phase separation to be the principle mechanism behind the self-association of shell protein in the presence of salts and macromolecular crowding. The shell protein PduBB’ interacts with the enzyme diol-dehydratase PduCDE and co-assemble into phase separated liquid droplets. The co-assembly of PduCDE and PduBB’ results in the enhancement of catalytic activity of the enzyme. The shell proteins that make up PduBB' (PduB and PduB') have contrasting self-assembly behavior. While N-terminal truncated PduB' has a high self-associating property and forms solid assemblies that separates out of solution, the longer component of the shell protein PduBM38L is more soluble and show least tendency to undergo phase separation. A combination of spectroscopic, imaging and biochemical techniques shows the relevance of divalent cation Mg2+ in providing stability to intact PduMCP. Together our results suggest a combination of protein-protein interactions and phase separation guiding the self-assembly of Pdu shell protein and enzyme in solution phase.
Article
Intrinsically disordered regions in proteins have been functionally linked to the protein–protein interactions and genesis of several membraneless organelles. Depending on their residual makeup, hydrophobicity or charge distribution they may remain in extended form or may assume certain conformations upon biding to a partner protein or peptide. The present work investigates the distribution and potential roles of disordered regions in the integral proteins of 1,2-propanediol utilization microcompartments. We use bioinformatics tools to identify the probable disordered regions in the shell proteins and enzyme of the 1,2-propanediol utilization microcompartment. Using a combination of computational modelling and biochemical techniques we elucidate the role of disordered terminal regions of a major shell protein and enzyme. Our findings throw light on the importance of disordered regions in the self-assembly, providing flexibility to shell protein and mediating its interaction with a native enzyme. Communicated by Ramaswamy H. Sarma
Article
Full-text available
Bacterial microcompartments (BMCs) are complex macromolecular assemblies composed of any outer protein shell that encases a specific metabolic pathway cargo. Recent research is now starting to unravel some of the processes that are involved in directing the enzyme cargo to the inside of the BMC. In particular, an article in this issue of J Bacteriol by N. W. Kennedy, C. E. Mills, C. H. Abrahamson, A. Archer, et al. (J Bacteriol 204:e00576-21, 2022, https://doi.org/10.1128/jb.00576-21) highlights the role played by the shell protein PduB in coordinating this internalization process.
Article
Full-text available
Spatial organization via encapsulation of enzymes within recombinant nanocompartments may increase efficiency in multienzyme cascades. Previously, we reported the encapsulation of single cargo proteins within nanocompartments in the heterologous host Escherichia coli. This was achieved by coexpression of the Salmonella enterica LT2 ethanolamine utilization bacterial microcompartment shell proteins EutS or EutSMNLK, with a signal sequence EutC1-19 cargo protein fusion. Optimization of this system, leading to the targeting of more than one cargo protein, requires an understanding of the encapsulation mechanism. In this work, we report that the signal sequence EutC1-19 targets cargo to the interior of nanocompartments via a hydrophobic interaction with a helix on shell protein EutS. We confirm that EutC1-19 does not interact with other Eut BMC shell proteins, EutMNLK. Furthermore, we show that a second signal sequence EutE1-21 interacts specifically with the same helix on EutS. Both signal sequences appear to compete for the same EutS helix to simultaneously colocalize two cargo proteins to the interior of recombinant nanocompartments. This work offers the first insights into signal sequence-shell protein interactions required for cargo sequestration within Eut BMCs. It also provides a basis for the future engineering of Eut nanocompartments as a platform for the potential colocalization of multienzyme cascades for synthetic biology applications.
Article
Full-text available
Bacterial microcompartments (MCPs) are complex organelles that consist of metabolic enzymes encapsulated within a protein shell. In this study, we investigate the function of the PduJ MCP shell protein. PduJ is 80% identical in amino acid sequence to PduA and both are major shell proteins of the 1,2-propanediol (1,2-PD) utilization (Pdu) MCP of Salmonella. Prior studies showed that PduA mediates the transport of 1,2-PD (the substrate) into the Pdu MCP. Surprisingly, however, results presented here establish that PduJ has no role 1,2-PD transport. The crystal structure revealed that PduJ was nearly identical to that of PduA and, hence, offered no explanation for their differential functions. Interestingly, however, when a pduJ gene was placed at the pduA chromosomal locus, the PduJ protein acquired a new function, the ability to mediate 1,2-PD transport into the Pdu MCP. To our knowledge, these are the first studies to show that that gene location can determine the function of a MCP shell protein. We propose that gene location dictates protein-protein interactions essential to the function of the MCP shell.
Article
Full-text available
Compartmentalization of designed metabolic pathways within protein based nanocompartments has the potential to increase reaction efficiency in multi-step biosynthetic reactions. We previously demonstrated proof-of-concept of this aim by targeting a functional enzyme to single cellular protein nanocompartments, which were formed upon recombinant expression of the Salmonella enterica LT2 ethanolamine utilization bacterial microcompartment shell proteins EutS or EutSMNLK in Escherichia coli. To optimize this system, increasing overall encapsulated enzyme reaction efficiency, factor(s) required for the production of more than one nanocompartment per cell must be identified. In this work we report that the cupin domain protein EutQ is required for assembly of more than one nanocompartment per cell. Overexpression of EutQ results in multiple nanocompartment assembly in our recombinant system. EutQ specifically interacts with the shell protein EutM in vitro via electrostatic interactions with the putative cytosolic face of EutM. These findings lead to the theory that EutQ could facilitate multiple nanocompartment biogenesis by serving as an assembly hub for shell proteins. This work offers insights into the biogenesis of Eut bacterial microcompartments, and also provides an improved platform for the production of protein based nanocompartments for targeted encapsulation of enzyme pathways.
Article
Full-text available
Bacterial microcompartments (BMCs) enhance the breakdown of metabolites such as 1,2-propanediol (1,2-PD) to propionic acid. The encapsulation of proteins within the BMC is mediated by the presence of targeting sequences. In an attempt to redesign the Pdu BMC into a 1,2-PD synthesising factory using glycerol as the starting material we added N-terminal targeting sequences to glycerol dehydrogenase, dihydroxyacetone kinase, methylglyoxal synthase and 1,2-propanediol oxidoreductase to allow their inclusion into an empty BMC. 1,2-PD producing strains containing the fused enzymes exhibit a 245% increase in product formation in comparison to un-tagged enzymes, irrespective of the presence of BMCs. Tagging of enzymes with targeting peptides results in the formation of dense protein aggregates within the cell that are shown by immuno-labelling to contain the vast majority of tagged proteins. It can therefore be concluded that these protein inclusions are metabolically active and facilitate the significant increase in product formation.
Article
Full-text available
Gammaproteobacteria are important gut microbes but only persist at low levels in the healthy gut. The ecology of Gammaproteobacteria in the gut environment is poorly understood. Here, we demonstrate that choline is an important growth substrate for representatives of Gammaproteobacteria. Using Proteus mirabilis as a model, we investigate the role of choline metabolism and demonstrate that the cutC gene, encoding a choline-trimethylamine lyase, is essential for choline degradation to trimethylamine by targeted mutagenesis of cutC and subsequent complementation experiments. P. mirabilis can rapidly utilise choline to enhance growth rate and cell yield in broth culture. Importantly, choline also enhances swarming-associated colony expansion of P. mirabilis under anaerobic conditions on solid surface. Comparative transcriptomics demonstrated that choline not only induces choline-trimethylamine lyase, but also genes encoding shell proteins for the formation of bacterial microcompartments. Subsequent analyses by transmission electron microscopy confirmed the presence of such novel microcompartments in cells cultivated in liquid broth and hyper-flagellated swarmer cells from solid medium. Together, our study reveals choline metabolism as an adaptation strategy for P. mirabilis and contributes to better understand the ecology of this bacterium in health and disease.
Article
Full-text available
Bacterial microcompartments (BMCs) are proteinaceous organelles encapsulating enzymes that catalyze sequential reactions of metabolic pathways. BMCs are phylogenetically widespread, however only a few BMCs are experimentally characterized. Among them are the carboxysomes and the propanediol and ethanolamine utilizing microcompartments, which play diverse metabolic and ecological roles. The substrate of a BMC is defined by its signature enzyme. In catabolic BMCs, this enzyme typically generates an aldehyde. Recently, it was shown that the most prevalent signature enzymes encoded among BMC loci are glycyl radical enzymes; yet little is known about the function of these BMCs. Here we characterize the glycyl radical enzyme-associated microcompartment (GRM) loci using a combination of bioinformatic analyses and active site and structural modeling, to show that the GRMs comprise five subtypes. We predict distinct functions for the GRMs, including the degradation of choline, propanediol, and fuculose-phosphate. This is the first family of BMCs for which identification of the signature enzyme is insufficient for predicting function. The distinct GRM functions are also reflected in differences in shell composition and apparently different assembly pathways. The GRMs are the counterparts of the B 12 -dependent propanediol and ethanolamine utilizing BMCs, which are frequently associated with virulence. This study provides a comprehensive foundation for experimental investigations on the diverse roles of GRMs. Understanding this plasticity of function within a single BMC family, including characterization of differences in permeability and assembly, can inform approaches to BMC bioengineering and the design of therapeutics.
Article
Bacterial microcompartments are naturally occurring subcellular organelles of bacteria, and serve as a promising scaffold for the organization of heterologous biosynthetic pathways. A critical element in the design of custom biosynthetic organelles is quantitative control over the loading of heterologous enzymes to the interior of the organelles. We demonstrate that the loading of heterologous proteins to the 1,2-propanediol utilization microcompartment of Salmonella enterica can be controlled using two strategies: by modulating the transcriptional activation of the microcompartment container, and by coordinating the expression of the microcompartment container and the heterologous cargo. These strategies allow general control over the loading of heterologous proteins localized by two different N-terminal targeting peptides, and represent an important step towards tuning the catalytic activity of bacterial microcompartments for increased biosynthetic productivity.
Article
Salmonella enterica forms polyhedral organelles involved in coenzyme B-12-dependent 1,2-propanediol degradation. These organelles are thought to consist of a proteinaceous shelf that encases coenzyme B-12-dependent diol dehydratase and perhaps other enzymes involved in 1,2-propanediol degradation. The function of these organelles is unknown, and no detailed studies of their structure have been reported. Genes needed for organelle formation and for 1,2-propanediol degradation are located at the 1,2-propanediol utilization (pdu) locus, but the specific genes involved in organelle formation have not been identified. Here, we show that the pduA gene encodes a shell protein required for the formation of polyhedral organelles involved in coenzyme B-12-dependent 1,2-propanediol degradation. A His(6)-PduA fusion protein was purified from a recombinant Escherichia coli strain and used for the preparation of polyclonal antibodies. The anti-PduA antibodies obtained were partially purified by a subtraction procedure and used to demonstrate that the PduA protein localized to the shell of the polyhedral organelles. In addition, electron microscopy studies established that strains with nonpolar pduA mutations were unable to form organelles. These results show that the pduA gene is essential for organelle formation and indicate that the PduA protein is a structural component of the shell of these organelles. Physiological studies of nonpolar pdu4 mutants were also conducted. Such mutants grew similarly to the wild-type strain at low concentrations of 1,2-propanediol but exhibited a period of interrupted growth in the presence of higher concentrations of this growth substrate. Growth tests also showed that a nonpolar pdu4 deletion mutant grew faster than the wild-type strain at low vitamin B-12 concentrations. These results suggest that the polyhedral organelles formed by S. enterica during growth on 1,2-propanediol are not involved in the concentration of 1,2-propanediol or coenzyme B-12, but are consistent with the hypothesis that these organelles moderate aldehyde production to minimize toxicity.