Pair your accounts.

Export articles to Mendeley

Get article recommendations from ACS based on references in your Mendeley library.

Pair your accounts.

Export articles to Mendeley

Get article recommendations from ACS based on references in your Mendeley library.

You’ve supercharged your research process with ACS and Mendeley!

STEP 1:
Click to create an ACS ID

Please note: If you switch to a different device, you may be asked to login again with only your ACS ID.

Please note: If you switch to a different device, you may be asked to login again with only your ACS ID.

Please note: If you switch to a different device, you may be asked to login again with only your ACS ID.

MENDELEY PAIRING EXPIRED
Your Mendeley pairing has expired. Please reconnect
ACS Publications. Most Trusted. Most Cited. Most Read
Structure of a Minimal α-Carboxysome-Derived Shell and Its Utility in Enzyme Stabilization
My Activity

Figure 1Loading Img
  • Open Access
Article

Structure of a Minimal α-Carboxysome-Derived Shell and Its Utility in Enzyme Stabilization
Click to copy article linkArticle link copied!

  • Yong Quan Tan
    Yong Quan Tan
    Department of Biochemistry, Yong Loo Lin School of Medicine, National University of Singapore (NUS), 8 Medical Drive, Singapore 117597
    NUS Synthetic Biology for Clinical and Technological Innovation, 28 Medical Drive, Singapore 117456
    Graduate School for Integrative Sciences and Engineering, NUS, Singapore 119077
  • Samson Ali
    Samson Ali
    Structural Biology Research Center, Graduate School of Science, Nagoya University, Furo-cho, Chikusa-ku, Nagoya 464-8601, Japan
    Research Institute for Interdisciplinary Science (RIIS), Okayama University, 3-1-1 Tsushima-naka, Kita-ku, Okayama 700-8530, Japan
    More by Samson Ali
  • Bo Xue
    Bo Xue
    Department of Biochemistry, Yong Loo Lin School of Medicine, National University of Singapore (NUS), 8 Medical Drive, Singapore 117597
    NUS Synthetic Biology for Clinical and Technological Innovation, 28 Medical Drive, Singapore 117456
    Synthetic Biology Translational Research Programme, Yong Loo Lin School of Medicine, National University of Singapore, 14 Medical Drive, Singapore 117599
    More by Bo Xue
  • Wei Zhe Teo
    Wei Zhe Teo
    Department of Biochemistry, Yong Loo Lin School of Medicine, National University of Singapore (NUS), 8 Medical Drive, Singapore 117597
    NUS Synthetic Biology for Clinical and Technological Innovation, 28 Medical Drive, Singapore 117456
    Synthetic Biology Translational Research Programme, Yong Loo Lin School of Medicine, National University of Singapore, 14 Medical Drive, Singapore 117599
    More by Wei Zhe Teo
  • Lay Hiang Ling
    Lay Hiang Ling
    Department of Biochemistry, Yong Loo Lin School of Medicine, National University of Singapore (NUS), 8 Medical Drive, Singapore 117597
    NUS Synthetic Biology for Clinical and Technological Innovation, 28 Medical Drive, Singapore 117456
    Graduate School for Integrative Sciences and Engineering, NUS, Singapore 119077
  • Maybelle Kho Go
    Maybelle Kho Go
    Department of Biochemistry, Yong Loo Lin School of Medicine, National University of Singapore (NUS), 8 Medical Drive, Singapore 117597
    NUS Synthetic Biology for Clinical and Technological Innovation, 28 Medical Drive, Singapore 117456
    Synthetic Biology Translational Research Programme, Yong Loo Lin School of Medicine, National University of Singapore, 14 Medical Drive, Singapore 117599
  • Hong Lv
    Hong Lv
    Shanghai Engineering Research Center of Industrial Microorganisms, Shanghai 200438, People’s Republic of China
    State Key Laboratory of Genetic Engineering, School of Life Science, Fudan University, Shanghai 200438, People’s Republic of China
    More by Hong Lv
  • Robert C. Robinson
    Robert C. Robinson
    Research Institute for Interdisciplinary Science (RIIS), Okayama University, 3-1-1 Tsushima-naka, Kita-ku, Okayama 700-8530, Japan
    School of Biomolecular Science and Engineering (BSE), Vidyasirimedhi Institute of Science and Technology (VISTEC), Rayong 21210, Thailand
  • Akihiro Narita
    Akihiro Narita
    Structural Biology Research Center, Graduate School of Science, Nagoya University, Furo-cho, Chikusa-ku, Nagoya 464-8601, Japan
  • Wen Shan Yew*
    Wen Shan Yew
    Department of Biochemistry, Yong Loo Lin School of Medicine, National University of Singapore (NUS), 8 Medical Drive, Singapore 117597
    NUS Synthetic Biology for Clinical and Technological Innovation, 28 Medical Drive, Singapore 117456
    Graduate School for Integrative Sciences and Engineering, NUS, Singapore 119077
    Synthetic Biology Translational Research Programme, Yong Loo Lin School of Medicine, National University of Singapore, 14 Medical Drive, Singapore 117599
    *Email: [email protected]
    More by Wen Shan Yew
Open PDFSupporting Information (1)

Biomacromolecules

Cite this: Biomacromolecules 2021, 22, 10, 4095–4109
Click to copy citationCitation copied!
https://doi.org/10.1021/acs.biomac.1c00533
Published August 12, 2021

Copyright © 2021 The Authors. Published by American Chemical Society. This publication is licensed under

CC-BY-NC-ND 4.0 .

Abstract

Click to copy section linkSection link copied!

Bacterial microcompartments are proteinaceous shells that encase specialized metabolic processes in bacteria. Recent advances in simplification of these intricate shells have encouraged bioengineering efforts. Here, we construct minimal shells derived from the Halothiobacillus neapolitanus α-carboxysome, which we term Cso-shell. Using cryogenic electron microscopy, the atomic-level structures of two shell forms were obtained, reinforcing notions of evolutionarily conserved features in bacterial microcompartment shell architecture. Encapsulation peptide sequences that facilitate loading of heterologous protein cargo within the shells were identified. We further provide a first demonstration in utilizing minimal bacterial microcompartment-derived shells for hosting heterologous enzymes. Cso-shells were found to stabilize enzymatic activities against heat shock, presence of methanol co-solvent, consecutive freeze–thawing, and alkaline environments. This study yields insights into α-carboxysome assembly and advances the utility of synthetic bacterial microcompartments as nanoreactors capable of stabilizing enzymes with varied properties and reaction chemistries.

This publication is licensed under

CC-BY-NC-ND 4.0 .
  • cc licence
  • by licence
  • nc licence
  • nd licence
Copyright © 2021 The Authors. Published by American Chemical Society

Introduction

Click to copy section linkSection link copied!

Bacterial microcompartments (BMCs) are protein shells found in some bacterial species that compartmentalize specialized biochemical reactions. (1) The carboxysome, found in cyanobacteria and some chemotrophic bacterial species, is the earliest known example of a BMC. (2,3) This semipermeable shell encapsulates ribulose-1,5-bisphosphate carboxylase (RuBisCO) and improves its catalytic efficiency by concentrating CO2 close to RuBisCO. (4,5) Carboxysomes are classified into two main groups depending on the class of RuBisCO that is encapsulated. α-Carboxysomes encase Form 1A RuBisCO that is found in α-cyanobacteria and chemoautotrophs, while β-carboxysomes house Form 1B RuBisCO that are observed in β-cyanobacteria. (2)
Numerous atomic-scale structures of the subunits that comprise BMC shells, along with that of three minimal BMC-derived shells (Table 1), have been reported. (6−10) Despite the diversity in the overall morphologies and functions of BMCs, the tertiary structures of the shell subunits are conserved. (11) The BMC-P (or BMV) protein is a minor but important subunit in a BMC shell that occupies the shell vertices as homopentamers, giving rise to the polyhedral morphology distinctive of BMCs. (12) The BMC-H domain protein is the stoichiometrically major subunit and assembles as a homohexamer. The BMC-P and BMC-H domains are structurally and evolutionarily distinct from each other and belong to two separate protein families, Pfam03319 and Pfam00936, respectively. (12) The BMC-T protein is formed by a tandem repeat of two BMC-H domains and assembles as a single or double stack of trimers, giving rise to a pseudo-hexameric appearance. BMC-H and BMC-T subunits constitute the shell facets and influence the diffusion of small molecules across the shell membrane. (1) Detailed understanding of the architecture of BMC components has been instrumental for BMC engineering toward synthetic biology applications. (13,14) Such endeavors include targeting heterologous cargo into the shell via the use of encapsulation peptides, which are peptides derived from cognate luminal proteins or by installation of peptide conjugation moieties into shell modules. (15−18)
Table 1. Summary of Structurally Elucidated Minimal BMC-Derived Shells Reported to Date
shellorganism of originnative shell functionshell structural models knownconstituent shell proteins identifiedreference
HO-BMCHaliangium ochraceumundeterminedT = 9BMC-P, BMC-H, BMC-TSutter et al. (7)
synthetic β-carboxysome shellsHalothece sp. PCC7428CO2 fixationT = 3, T = 4BMC-P, BMC-HSutter et al. (6)
GRM2 BMC-derived particlesKlebsiella pneumoniaecholine catabolismT = 4BMC-P, BMC-TKalnins et al. (10)
Cso-shellH. neaCO2 fixationT = 3, T = 4BMC-P, BMC-Hthis study
The α-carboxysome from Halothiobacillus neapolitanus (henceforth referred to as H. nea) has been recombinantly produced in Escherichia coli by transplantation of the α-carboxysome operon (cso) (Figure 1). (19,20) Genes found in the operon include three BMC-H paralogs (cso1ABC), one BMC-T (csoS1D) and two BMC-P paralogs (csoS4AB), along with those that encode the RuBisCO (cbbLS), carbonic anhydrase (csoSCA) and csoS2, an intrinsically disordered protein that serves as a crucial scaffold for α-carboxysome assembly. (21) A native-like α-carboxysome, expressed from six shell proteins and CsoS2, has been recently repurposed to host a hydrogenase complex for microbial hydrogen production, illustrating the bioengineering potential of the shell. (22) We reasoned that building a minimal shell derived from α-carboxysome components can yield further insights into α-carboxysome biogenesis, while improving their tractability for widespread adoption in biotechnology.

Figure 1

Figure 1. H. nea α-carboxysome operon (cso). The dotted line indicates 10 genes between cbbO and csoS1D that are unlikely to be associated with shell assembly or function. Gene lengths and intervening distances are not drawn to scale.

We report that protein shells resembling icosahedral capsids can assemble using two subunits, representing a minimal BMC-derived shell which we term Cso-shell. Sequences derived from CsoS2 were found to mediate cargo protein encapsulation, with one variant likely displaying the highest cargo loading efficacy for a minimal BMC-derived shell to date. Atomic-scale models of the Cso-shells were obtained using cryogenic electron microscopy, revealing two shell forms. Despite the evolutionary divergence of the α-carboxysome from other BMCs, the Cso-shell shares key features with other minimal BMC-derived shells, substantiating the idea that principles of BMC shell assembly are highly conserved. (1,6,7,10,23) Hosting of cargo enzymes within the shell generally increased tolerance of the enzymes toward heat shock, methanolic buffers, consecutive freeze–thaw cycles and alkaline environments. We propose that the Cso-shell can serve as a platform for understanding α-carboxysome assembly. Additionally, the enzyme-stabilizing effect of the Cso-shell represents an advancement in harnessing BMC-based technology for catalytic applications. (24)

Materials and Methods

Click to copy section linkSection link copied!

Bacterial Strains and Culture

E. coli Acella(DE3) (EdgeBio) cells were used for all molecular cloning and all protein expression experiments except for expression of recombinant LacZ (either free enzyme or encapsulated within shells). BL21(DE3)-Gold Δlac (Addgene accession 99247) was used for expression of constructs containing His6-LacZ-S2-CTP to prevent co-purification of endogenous LacZ. (25) The cells were grown in Terrific Broth (BioBasic) supplemented with the appropriate antibiotics (kanamycin or streptomycin) at 50 μg/mL.

Golden Gate Assembly of Plasmids

The Golden Gate one-pot plasmid assembly largely follows previously published protocols, with slight modifications. (26) All molecular cloning reagents were purchased from NEB. For the insertion of one to three fragments, in a reaction pot was prepared 1 μL of T4 ligase buffer, 0.2 μL of 10x purified bovine serum albumin (BSA), 5 U of BsaI or Esp3I, 1 U of T4 ligase, 15 ng of destination plasmid, 1 to 3 μL of insert(s) and topped up to 10 μL with water. The reaction pot was subject to a 37 to 18 °C thermocycling process with 5 min incubation at each step for 40 cycles, followed by a 55 °C step for 15 min to digest unassembled plasmids while inhibiting ligation so as to reduce the number of colonies harboring the original destination vector.
Codon-optimized BMC genes were synthesized (BioBasic) while promoters and terminator parts were amplified from various templates as PCR products and cloned into Level 0 plasmids (termed HcKan). Modified Level 1 and 2 plasmids (Figure S1) were created using NEB HiFi assembly.

Sequence Alignment and Bioinformatics

CsoS2 sequences were aligned with Clustal Omega and output alignment file was prepared with JalView 2. (27,28) Accession numbers for the sequences used in alignment are detailed in Table S3.

Protein Purification and Densitometric Analysis

Cells were cultured in 500 mL of terrific broth (TB) supplemented with the appropriate antibiotic and shaken at 37 °C until the OD600 value of the culture reached approximately 0.6–1.0. The cultures were then cooled to 25 °C and isopropyl β-d-1-thiogalactopyranoside (IPTG) was added to 100 μM for protein induction. The cells were shaken at 25 °C in 2 L baffled flasks at 150 rpm for 22 to 24 h before harvesting. Lysis was carried out using the M-110P microfluidizer (Microfluidics) at 15 000 psi for three passes in approximately 100 mL of lysis buffer (Tris·HCl 100 mM, NaCl 150 mM, pH 8.0). To the lysis mixture was also added 0.1 mM of phenylmethylsulfonyl fluoride (PMSF) protease inhibitor. The lysate was spun twice at 20 000g for 20 min each time. The clarified lysate was loaded at 1 mL/min linear flow rate onto either a StrepTrap HP 5 mL column (GE Life Sciences) for SII-tagged proteins or a HisPur cobalt 5 mL column (Thermo Scientific) for His6 tagged proteins. Purification was performed using an ÄKTA FPLC with 12 column volumes (CV) of washing with binding buffer (StrepTrap: Tris·HCl 100 mM, NaCl 150 mM, pH 8.0; HisPur: Tris·HCl 100 mM, NaCl 300 mM, imidazole 5 mM, pH 8.0) and 6 CVs of elution with the elution buffer (StrepTrap: binding buffer with 2.5 mM desthiobiotin; HisPur: binding buffer with 200 mM imidazole).
Shell samples were subject to anion-exchange chromatography following Strep-Tactin purification. Samples were dialyzed into ion exchange Buffer A (Tris·HCl 50 mM, pH 8.0) and loaded onto 10 mL of Q Sepharose HP resin in an XK 16/20 column (GE Life Sciences) at 1 mL/min. Buffer flow was kept at 1 mL/min and a two-step gradient protocol consisting of 0 to 50% ion exchange Buffer B (Tris·HCl 50 mM, NaCl 1.0 M, pH 8.0) over 15 CV and 60 to 100% Buffer B over 1.5 CV was used for elution. Shell-containing samples were concentrated using an Amicon centrifugal filter with 100 kDa nominal molecular weight limit (NMWL). Shells co-expressed with His6-APEX2-S2-CTP or His6-LacZ-S2-CTP were incubated with a 0.5 mL slurry of HisPur cobalt resin for 1 h at room temperature with gentle shaking to remove His6-tagged proteins not encapsulated within shells. The mixture was spun down at 20 000g for 5 min, and the supernatant was used for further analysis.
Proteins were analyzed using 13% stacking sodium dodecyl sulfate-polyacrylamide gel electrophoresis (SDS-PAGE) gels and stained using InstantBlue (Expedeon). Densitometric analysis was performed using the Bio-Rad image lab software in accordance with previous reports on BMC cargo quantification. (17,29) Background subtraction was performed and peak areas corresponding to bands of interest were used for quantification. Absolute protein concentrations were measured using a DeNovix spectrophotometer using calculated molar attenuation coefficients at 280 nm (ε280). The calculated ε280 of T = 3 shells, taken as the summation of the ε280 of its individual components, was 1 588 200 M–1·cm–1. The ε280 of the T = 4 shell was calculated to be 1 677 600 M–1·cm–1. The small difference in calculated ε280 values between both shell types is due to the low ε280 of CsoS1A (1490 M–1·cm–1). Most of the ε280 contribution comes from CsoS4A (23 490 M–1·cm–1), which has the same number of copies in both shell types. Concentrations of GFP were determined by fluorescence emission at 530 nm. Known concentrations of GFP were used to create calibration curves.

Immunoblotting

Proteins were transferred from SDS-PAGE gels to poly(vinylidene fluoride) (PVDF) membranes via the Bio-Rad Trans-Blot system. After transfer, membranes were blocked with phosphate-buffered saline-Tween 20 (PBS-T, Tween 20 at 0.05 v/v %) containing 3% w/v bovine serum albumin (BSA) for 1 h at room temperature with gentle shaking. Antibodies were prepared in blocking buffer at 1:1000 dilution, added to the blots, and incubated for 1 h with gentle shaking. For detection of GFP epitope, a GFP antibody conjugated with HRP (GF28R, Invitrogen) was used. For detection of His6, THE His Tag antibody conjugated with HRP (Genscript) was used. Blots were washed three times with PBS-T before chemiluminescent detection using the Amersham ECL Prime Western Blotting Detection Reagent (GE Healthcare Life Sciences).

Electron Microscopy

Formvar/carbon-coated copper grids (Electron Microscopy Science) were subject to glow discharge before 5 μL of purified protein samples (diluted to A280nm ∼ 0.05) were mounted for 60 s before the droplets were removed with filter paper. The grids were then negatively stained by adding a 5 μL droplet of 2.5% gadolinium(III) acetate, incubating for 90 s, and similarly blotted. Grids were imaged using a JEOL JEM-1220 TEM.

Shell Particle Size and Stability Measurements

Particle size distribution was determined by dynamic light scattering (DLS) using the UNcle instrument (Unchained labs). Samples were diluted to 1 mg/mL in TBS-50/350 pH 8.0 (Tris·HCl 50 mM, NaCl 350 mM, pH 8.0), unless otherwise specified, and spun at 20 000g for 5 min to remove aggregates prior to measurement. Care was taken to use the top-most supernatant for analysis. To the mini cuvette was added 9 μL of sample. All DLS measurements were done in triplicate and performed at 20 °C unless stated otherwise. Analysis of particle size distribution was done using the UNcle analysis software.
For shell stability measurements at various temperatures, shell samples were aliquoted into thin-walled PCR tubes and subject to temperatures ranging from 20–80 °C in 10 °C increments (indicated in Figure S8A) for 15 min in the UNcle instrument. At the end of the 15 min incubation, DLS spectra were taken.
For shell stability measurements in various buffer conditions, TBS-50/350 pH 8.0 buffers containing 10 and 20% (v/v) methanol were freshly prepared from stock solutions of Tris·HCl 1.0 M pH 8.0, NaCl 5.0 M, and 99.8% methanol (ACS reagent grade, Sigma) and used within the day of preparation. Due to the heat of mixing generated when methanol is mixed with water, methanol-containing buffers were allowed to equilibrate back to room temperature for at least 1 h after preparation. For making buffers at various pH, the following components were used at 50 mM for the pH range indicated: glycine·HCl for pH 2–4; 4-morpholineethansulfonic acid (MES) sodium salt for pH 5–7; Tris·HCl for pH 8–9; N-cyclohexyl-3-aminopropanesulfonic acid (CAPS) sodium salt for pH 10–11; arginine·HCl for pH 12–13. All buffers contained 350 mM NaCl. Shells were incubated in the above-mentioned buffers for 15 min to allow time for possible shell dissociation/protein denaturation before particle size measurement.
For freeze–thaw stability, shell samples in TBS-50/350 pH 8.0 were aliquoted into thin-walled PCR tubes and flash-frozen in liquid N2. Samples were thawed at room temperature until no ice crystals were seen by standing for 15 min before re-freezing.

Isothermal Titration Calorimetry

Isothermal titration calorimetry measurements were carried out using the MicroCal PEAQ-ITC. Receptor samples (e.g., free CsoS1A or an equivalent amount of CsoS1A within Cso-shells) were loaded in the cell while ligand samples (S2-C variants) were loaded in the syringe. Protein samples were purified via Strep-Tactin by installing the SII tag at the C-terminus of CsoS1A and at the N-termini of S2-C variants. S2-CTP was chemically synthesized (BioBasic). Protein/peptide samples were dialyzed to HEPES 20 mM pH 7.4 containing 20, 60, or 200 mM NaCl. Samples were spun at 20 000g for 5 min to remove aggregates prior to measurement. Receptor concentrations were adjusted to 10 μM while ligands were adjusted to 100 μM, with the exception of experiments involving S2-CTP. For S2-CTP, receptor concentration had to be increased to 50 μM and ligand to 500 μM before binding could be detected. The stirring speed was set to maximum and 20 injections were performed at 25.0 °C. Binding analysis was carried out using the MicroCal PEAQ-ITC Analysis software using one set of binding sites as the model.

Enzyme Steady-State Kinetics Assays

For APEX2, all reagents were made to the appropriate working concentrations using TBS-50/350 pH 8.0. Working solutions of guaiacol and H2O2, both at 10 mM, were prepared on the day of assay. The guaiacol solution was shaken vigorously at 30 °C to ensure complete dissolution before equilibration back to room temperature. The assay concentration for APEX2 was 10 nM, H2O2 1 mM, and that for guaiacol ranged from 0.20 to 2.0 mM. The total reaction volume was 200 μL. Reactions were monitored by the absorbance of tetraguaiacol formed at 470 nm using a BioTek Synergy HT microplate reader. The rate of tetraguaicol formation was found to be constant until 90 s. This time point was taken for initial rate, V0, measurement. Kinetic constants were obtained using nonlinear least-square Michaelis–Menten fitting in GraphPad Prism.
For steady-state kinetics of LacZ, all reagents were made to the appropriate working concentrations using TBS-50/350 pH 8.0 supplemented with 1 mM MgCl2. Working solutions of ONPG at 10 mM were prepared from a 50 mM stock solution in DMSO on the day of assay. The assay concentration for LacZ was 10 nM, and that for ONPG (ortho-nitrophenyl-β-galactoside) ranged from 0.050 to 1.5 mM. The total reaction volume was 100 μL. Hydrolysis of ONPG was tracked by absorbance measurements at 405 nm. Rate of production formation was found to be constant until 60 s, and this time point was taken for V0 measurement.

Enzyme/Enzyme-Shell Activity and Stability Assays

All enzyme activity measurements were performed at ambient temperature (23 °C) and working concentrations of enzymes were 10 nM. Measurements were performed in triplicate. Enzyme activity measurements were determined as initial rates of product formation using saturating substrate concentrations (i.e., near Vmax). For APEX2, this was 1.4 mM guaiacol and 1 mM H2O2. For LacZ, this was 1.5 mM ONPG.
For heat shock assays, enzyme/enzyme-shell samples were aliquoted into thin-walled PCR tubes and subject to the elevated temperatures indicated (Figure 6B) in a thermocycler for 15 min. Following incubation, samples were cooled to 20 °C and equilibrated back to ambient temperature for 15 min before assay.
For stability measurement in buffers containing methanol and under various pH conditions, enzyme/enzyme-shell samples were dialyzed to the various buffers as described in the particle size measurement section. Solutions were allowed to stand for at least 15 min before assay to allow time for possible protein denaturation.
For freeze–thaw stability, enzyme/enzyme-shell samples were aliquoted in thin-walled PCR tubes and flash-frozen in liquid N2. Samples were thawed at room temperature until no ice crystals were seen by standing for 15 min before re-freezing or assay.

Results and Discussion

Click to copy section linkSection link copied!

Formation of Shell-like Particles from Cso-Shell Components

We sought to derive minimal shells from the cso operon using only one member from each shell subunit class (BMC-P/-H/-T). In structurally characterized BMC-derived shells reported to date (Table 1), BMC-P and BMC-H have been essential, while BMC-T has only been unambiguously identified in HO-BMC, indicating the BMC-T subunit is sometimes present in such shells. Our strategy involved combining subunits as PTH or PH, where P represents CsoS4A or CsoS4B (BMC-P), T represents CsoS1D (BMC-T), and H represents CsoS1A, CsoS1B, or CsoS1C (BMC-H), for a total of 12 combinations. A recently reported affinity purification strategy involving the installation of the Strep-tag II (SII) to the C-terminus of the BMC-P subunit was used to expedite purification of potential shells. (17) To facilitate combinatorial synthesis of BMC pathways, we adopted an established hierarchical Golden Gate modular cloning approach (Figures 2 and S1). (26) Briefly, promoter, open reading frame (ORF) and terminator parts are maintained in Level 0 plasmids. These parts are strung together as promoter-ORF-terminator transcriptional units in Level 1 plasmids. Transcriptional units are then combined into a contiguous pathway in a Level 2 acceptor plasmid.

Figure 2

Figure 2. Identifying combinations that produce minimal α-carboxysome-derived shells. (A) Schematic of the combinatorial assembly process used to construct shell-expressing pathways. (B–E) Transmission electron microscopy (TEM) micrographs of combinations that produced capsid-like structures, (B) Cso-PATHA, (C) Cso-PAHA, (D) Cso-PATHC, and (E) Cso-PAHC. Scale bars (bottom right) represent 50 nm. (F, G) SDS-PAGE analysis of shell-containing samples following anion-exchange purification of (F) Cso-PATHA and (G) Cso-PAHA. The protein ladder is indicated by L in kDa.

Controlling the stoichiometry of BMC components is key to their successful assembly. (6,17) As a general guideline for expression of synthetic BMCs, the target relative expression levels for BMC-P, BMC-T, and BMC-H are low, medium, and high, respectively. (16,30) Correspondingly, in our modular system, a low-expression-strength promoter (BBa_J23114 from the Anderson Collection, Table S1) was assigned to P, a medium-strength promoter (BBa_J23105) to T, and the strong T7 promoter (PT7) to H. (31,32) The BBa_J23114 and BBa_J23105 promoters are constitutively active while PT7 is regulated by the LacI repressor protein to suppress high-level expression of recombinant proteins during the initial stage of cell growth, which is deleterious to the host cell. (33) Upon reaching the mid-exponential phase of cell growth, isopropyl β-d-1-thiogalactopyranoside (IPTG) was added to relieve the repression, enabling high-level protein expression of H.
Assembled transcriptional units were combined into pathways (Figure 2A). The 12 pathway combinations (Table S2) were expressed, affinity purified, and visualized by transmission electron microscopy (TEM). Four combinations produced shell-like particles, namely, CsoS4A-SII in combination with either CsoS1A or CsoS1C, with and without the co-expression of CsoS1D (i.e., Cso-PATHA, PAHA, PATHC, PATHC). Particles from the four samples, which can be described as capsid-like structures, were indistinguishable from each other (Figure 2B–E). These particles also closely resembled the synthetic β-carboxysome and GRM2 BMC-derived particles. (6,10) We noted that the sequence of CsoS1A is highly similar to CsoS1C, differing only by one residue near the N-terminus and another near the C-terminus (Figure S2A). Unsurprisingly, structural models of CsoS1A and CsoS1C are virtually identical (Figure S2B). As the two differing residues are in flexible regions unlikely to be important for quaternary assembly, we postulated that shells comprised CsoS4A in combination with either CsoS1A or CsoS1C should be equivalent. We thus focused on Cso-PATHA and Cso-PAHA for the remainder of this study. To determine the components necessary for shell formation, pathway combinations expressing CsoS4A-SII with CsoS1D and CsoS1D with CsoS1A-SII were created and similarly purified. No nanoscale structures could be seen. Thus, CsoS4A and CsoS1A were found to be necessary and sufficient for the formation of shell-like particles observed. We term these particles Cso-shells.
Affinity purified Cso-PATHA and Cso-PAHA shells were subject to anion-exchange chromatography to separate shells from unincorporated CsoS4A-SII. (17) Two well-separated peaks were seen in the chromatogram, corresponding to elution centered at 0.18 and 0.35 M NaCl (Figure S2C). Samples from both peaks were analyzed by TEM and shells were seen only in the 0.35 M NaCl fraction. Based on the observation that the elution peak of monomeric CsoS4A-SII was centered at 0.18 M NaCl, we reasoned that the 0.18 M NaCl peak for shell samples largely corresponded to CsoS4A-SII not incorporated into the shell. SDS-PAGE analysis of anion-exchange purified Cso-PATHA and Cso-PAHA shells showed the presence of a band that corresponds to both CsoS4A-SII and CsoS1A, which have calculated molecular masses of 10.0 kDa and could not be separated (Figure 2F,G). No band corresponding to CsoS1D (23.5 kDa) was detected in the Cso-PATHA sample, indicating CsoS1D was absent or at very low occupancy within the shell.

Identifying Encapsulation Peptide Sequences for Cso-Shell Cargo Loading

We endeavored to repurpose the Cso-shell for encapsulating protein cargo using encapsulation peptides. For the purposes of engineering BMC shells, encapsulation peptides are sequences that interact with the shell subunits, facilitating encapsulation of protein cargo fused to such peptides before the shell is fully assembled. (16,18,30) Bioinformatic analysis on the scaffolding protein, CsoS2, suggests it can be divided into three distinct regions: the N-terminal region (S2-N), middle region (S2-M), and C-terminal region (S2-C) (Figure 3A). (20) S2-C has also been postulated to initiate carboxysome assembly by recruiting CsoS1 proteins. (20,21,34) Hence, encapsulation peptide sequences may be derived from S2-C. (22) Close inspection of S2-C shows it contains three homologous repeat sequences (R1–R3) and a highly conserved 30 residue C-terminal peptide (CTP) (Figure S3). (20) To probe the ability of S2-C-derived peptides to mediate encapsulation in the Cso-shell, full-length S2-C and truncated sequences─S2-C-ΔR1, S2-C-ΔR1-2, and S2-CTP were fused to the C-terminus of a monomeric green fluorescent protein (GFP). An E. coli N-terminal degradation tag, UmuD (1−40) (first 40 residues of UmuD), which is recognized for proteolysis by the endogenous ClpXP protease, was installed on the N-terminus of GFP-S2-C variants. (35,36) Based on known structures of BMC-derived shells, the shell subunits are tightly interdigitated and impermeable toward biological macromolecules, such as proteases (Figure 3B). (7,17,37) On the other hand, UmuD1-40-tagged proteins that are not encapsulated or only interact with the shell periphery would be exposed to and degraded by ClpXP. Hence, this assay served to identify successfully encapsulated cargo proteins within intact shells.

Figure 3

Figure 3. Identifying S2-CR sequences for mediating cargo encapsulation within the Cso-shell. (A) CsoS2 can be split into three distinct regions based on sequence homology. Regions are not drawn to scale. (B) Schematic of the encapsulation-degradation assay. Encapsulation of GFP cargo within the shell prevents its degradation by the ClpXP protease. The blue and yellow blocks represent CsoS1A (BMC-H) and CsoS4A (BMC-P), respectively. (C) Pathway design for co-expressing UmuD (1−40)-GFP-S2-Cx (S2-Cx represents S2-C variants) with the Cso-PAHA shells. (D) Western blot detection of GFP in purified shell samples. The identities of the proteins in the ladder (L) lane are: His6-GFP-SpyTag (29.7 kDa), His6-GFP-SpyCatcher (38.2 kDa), and His6-GFP-SpyCatcher-SpyTag-GFP (67.9 kDa). Lanes 1–5 are Cso-PAHA co-expressed with the UmuD (1−40)-GFP variant indicated below the lanes. (E) DLS analysis of shells co-expressed with various cargo. The numbers to the right of the spectra correspond to the samples loaded in (D). The baseline is vertically displaced by 0.2 for each subsequent spectrum so that all peaks can be seen in one graph.

The UmuD (1−40)-GFP-S2-C variants were co-expressed with Cso-PAHA shells (Figure 3C). To account for stochastic encapsulation, a strain expressing UmuD (1−40)-GFP with shells was included. Purified shells were subject to Western blot analysis for detection of GFP (Figure 3D). Coomassie blue densitometric analysis of the bands corresponding to shell proteins was within ±0.4% of the mean value (Figure S4A). Hence, the quantity of shells loaded into each lane was approximately the same and the abundance of each UmuD (1−40)-GFP-S2-C variant, measured via immunoblotting, could be compared with each other (Table 2). S2-CTP mediated the highest level of cargo encapsulation, followed by S2-C-ΔR1-2. A faint band could be seen for UmuD (1−40)-GFP-S2-C-ΔR1 while both UmuD (1−40)-GFP and UmuD (1−40)-GFP-S2-C could not be detected, signifying very low occupancy or absence of encapsulated GFP. Encapsulation of the various cargo had no significant impact on shell size distribution (Figure 3E) and average hydrodynamic diameters (Table 2), as determined by dynamic light scattering (DLS). There was also no discernable difference in morphology among shell samples (Figure S4B–F).
Table 2. Densitometric Analyses of UmuD (1−40)-GFP Tagged with S2-C Variants, along with Average Hydrodynamic Diameters of the Shells Co-expressed with These Cargo Molecules
cargo (UmuD (1−40)-GFP-)cargo molecular mass (kDa)normalized peak area (immunoblot densitometry)average hydrodynamic diameter (nm)
No tag31.5not detected24.4
S2-C59.9not detected24.6
S2-C-ΔR150.90.03524.4
S2-C-ΔR1-242.20.41923.7
S2-CTP34.41.00024.0

Atomic-Scale Models of the Cso-Shell Reinforce Notions on Plasticity and Generalizability of BMC Shell Architecture

To better understand the molecular principles behind Cso-shell assembly and cargo encapsulation, cryogenic electron microscopy was used to obtain atomic-scale models of shells expressed from Cso-PAHA and Cso-PATHA. We wished to visualize GFP-S2-CTP density within the shell without possible interference from the UmuD (1−40) tag and thus created Cso-PAHA + GFP-S2-CTP (Figure 4A). For Cso-PATHA and Cso-PAHA, shells with icosahedral triangulation numbers T = 3 were seen, and maps were obtained at resolutions of 3.35 and 3.14 Å, respectively. For Cso-PAHA + GFP-S2-CTP, T = 3 and T = 4 shell forms were observed (Figure S5A), at proportions of 14 and 86%, respectively. Maps were obtained at resolutions of 3.24 Å for T = 3 and 2.90 Å for T = 4. Known structural models of H. nea CsoS1A and CsoS4A were used for model fitting. (8,38) We also solved the structure of H. nea CsoS1D at 2.70 Å resolution using X-ray crystallography. H. nea CsoS1D assembles as a stacked dimer of trimers with pseudo-hexameric symmetry (Figure S5B), similar to a homolog from Prochlorococcus marinus MED4. (39) However, in the Cso-PATHA model, no double-stacking densities were observed. The densities corresponding to hexameric units in the Cso-PATHA map could be modeled unequivocally with CsoS1A hexamers but not with a single layer of CsoS1D trimers. This observation, along with the absence of a detectable CsoS1D band in SDS-PAGE analysis of shells purified from Cso-PATHA (Figure 2F), led us to conclude that CsoS1D was unlikely to be incorporated within these shells.

Figure 4

Figure 4. Atomic-scale models of the minimal α-carboxysome shells. (A) Operon schematic of Cso-PAHA + GFP-S2-CTP. (B, C) Surface representations of the shells, with the CsoS1A colored blue and CsoS4A in yellow. The blue and red patches over yellow indicate N- and C-termini of a CsoS4A monomer, and the cyan and red patches over blue indicate those of a CsoS1A monomer.

Unfortunately, cargo density was not resolved within Cso-PAHA + GFP-S2-CTP shells. We surmised that the cargo was not rigidly bound within the shell, resulting in nonuniform luminal densities that prevented resolution. Nonetheless, in light of a theoretical study on the α-carboxysome that suggests encapsulated scaffold molecules may increase shell size, it is conceivable that the T = 4 shell, seen only for Cso-PAHA + GFP-S2-CTP, could result from cargo encapsulation. (40) As the densities for all T = 3 shell models were largely identical, we focused on the model produced by Cso-PAHA + GFP-S2-CTP for further refinement. The T = 3 shell contains 12 homopentamers of CsoS4A and 20 homohexamers of CsoS1A (Figure 4B) while the T = 4 shell contains 12 homopentamers and 30 homohexamers (Figure 4C). Both shell forms are similar in other aspects.
In both T = 3 and 4 shell forms, the concave sides of CsoS1A and CsoS4A, wherein the N- and C-termini reside, face away from the shell (Figure 4B,C). Compared with their structures in isolation, both CsoS1A and CsoS4A expand slightly in the shell (Figure S5C,D and Table S6). Expansion of CsoS4A brings its V11 and T13 residues closer to a neighboring CsoS1A hexamer (Figure 5A). Furthermore, the C-terminus of CsoS4A flips inward and anchors itself into a pocket formed by two adjacent CsoS4A monomers. This conformational change is stabilized by π–π interactions between W80 and W49 and cation−π interactions between W80 and K8 from an adjacent CsoS4A monomer. Conformational changes for the CsoS1A hexamer following shell incorporation include slight outward movement of the helices sandwiching the central β-sheet in the CsoS1A chain, which largely accounts for the hexamer expansion and repositioning of several solvent-exposed side chains (Figure 5B).

Figure 5

Figure 5. Close examination of the interfacial contacts in Cso-shells reinforcing notions of conserved features in BMC shell assembly. (A) CsoS4A conformations pre-shell (gray) and post-shell integration (yellow). The dashed arrow indicates movement of CsoS4A’s C-terminal loop. W80 is highlighted with dark gray (pre-integration) and orange (post-integration). (B) CsoS1A hexamers (gray) become expanded after shell incorporation (blue). Arrow indicates movement of the helices. (C) Pentamer–hexamer interface, T = 3. (D) Pentamer–hexamer interface, T = 4. (E) Hexamer–hexamer interface, T = 3. (F) Hexamer–hexamer interface, T = 4. For clarity, only half of the polar interactions left to the pseudo-2-fold axis (orange oval) at the hexamer–hexamer interface are indicated in (E) and (F). For (C)–(F), CsoS4A is colored yellow, while CsoS1A chains are colored blue and green. Dotted lines represent interatomic distances in Å.

Examination of the shell subunit interface reveals that hydrophobic patches contribute to interface assembly through shape complementarity while hydrogen bonds and salt bridges dictate spatial orientations of subunits (Figure 5C–F). The interfaces between CsoS1A and CsoS4A are conserved in both shell forms. The most prominent interfacial residue on CsoS4A is D48, which forms a salt bridge with R83 on CsoS1A (Figure 5C,D). In BMC-P orthologs, this position is generally found to be occupied by an acidic residue, denoting the importance of this polar engagement to the pentamer–hexamer interface. (1,7) Two interfacial motifs on CsoS1A, KAA (residues 29–31) and RxH (residues 83–85), constitute a large portion of both the pentamer–hexamer and the hexamer–hexamer interfaces. Key polar interactions are provided through K29 of the KAA motif and R83 of RxH. These motifs are known for their critical roles in BMC shells. (7,41) The side chain of R83 adopts two alternative conformations in the T = 3 shell; in the T = 4 shell, it takes a single, more engaging pose (Figures 5E,F and S5E,F). The morphological plasticity frequently observed in both native and synthetic BMC shells may be partly attributed to residues with flexible side chains in the KAA and RxH motifs that permit subunits to adjoin with differing interfacial orientations. (6,7,10)
Phylogenetic analyses of BMC-H and BMC-P domain proteins indicate that the α-carboxysomes shell proteins are evolutionarily divergent from β-carboxysomes and other BMCs, such as the metabolosomes that encase catabolic reactions. (23,42) Despite carrying out essentially the same function as α-carboxysomes, β-carboxysomes are considered to be more closely related to metabolosomes in terms of shell protein sequences and assembly mechanism. (1,23) Notwithstanding the distinct phylogenetic distance between the α-carboxysome and other BMCs, there are key features shared between the Cso-shell and other BMC-derived shells. The overall structural models of both Cso-shell forms are highly similar to that of the synthetic β-carboxysome, for which T = 3 and 4 forms are known, and also to the GRM2 BMC-derived particle, for which the T = 4 structure has been resolved. (6,10) In all BMC-derived shell models reported so far, the convex faces of the BMC-H and BMC-P subunits always face the shell lumen, and the KAA and RxH motifs are vital to interfacial contacts (Figure S5G). (6,7,10) The striking similarities observed among shells derived from evolutionarily and functionally divergent BMCs substantiate the idea that principles of BMC shell architecture are generalizable. (1,7)

Sequence Length of S2-C Peptides Appears to Influence Encapsulation Efficacies of S2-C Peptides In Vivo

The average copy numbers of UmuD (1−40)-GFP-S2-C variant encapsulated per shell could be quantified using the shell structures and GFP fluorescence. (17) Based on cryogenic electron microscopy observations, for all calculations involving shell molecular masses we presumed all shells co-expressed with cargo to be a mixture of T = 3 and 4 forms. As the proportions of shell forms may vary among samples, two values, calculated by presuming all shells in a sample were either T = 3 or 4, are provided. In agreement with results from immunoblotting, S2-CTP was the most efficacious encapsulation peptide, followed by S2-C-ΔR1-2 and S2-C-ΔR1 (Table 3 and Figure S6). Virtually no signal was detected for UmuD (1−40)-GFP-S2-C and untagged UmuD (1−40)-GFP.
Table 3. Quantifying the Average Number of Encapsulated UmuD (1−40)-GFP Cargo Per Shell as Mediated by S2-C-Derived Peptidesa
 average number of UmuD (1−40)-GFP cargo per shell  
S2-C variantT = 3T = 4number of residuesestimated hydrodynamic radius (Å)
no tag0.0230.024N.A.N.A.
S2-C0.110.1126642.7
S2-C-ΔR10.480.5018035.0
S2-C-ΔR1-22.93.09725.6
S2-CTP7.78.03014.1
a

As the proportions of the shell forms are unknown, values calculated by presuming all shells are either T = 3 or 4 are given. The estimated hydrodynamic radius of the peptides may be a factor contributing to the encapsulation efficacies observed.

While we were unable to obtain atomic-scale models of how S2-CTP interacts with the shell lumen, it is still of interest to consider possible factors that contribute to the observed encapsulation efficacies of the S2-C-derived peptides. Savage et. al. recently hypothesized that S2-C interacts with CsoS1 proteins during α-carboxysome biogenesis, and formation of a protein condensate arising from high local concentrations of CsoS2 lowers the ionic strength of the immediate environment, strengthening electrostatic interactions between α-carboxysome components. (20,21,34) However, no direct evidence of S2-C interacting with CsoS1A has been reported. This prompted us to investigate putative interactions of S2-C-derived peptides with CsoS1A using isothermal titration calorimetry (ITC). In HEPES 20/200 (HEPES 20 mM pH 7.4, NaCl 200 mM), which emulates physiological salt concentration, no binding was detected between 100 μM of S2-C peptides separately titrated against 10 μM of isolated CsoS1A. (43) When solution ionic strength was lowered by decreasing NaCl concentration to 60 mM, binding was detected for all S2-C peptides (Figure 6A). The standard binding free energies, Δ, corresponded to a dissociation constant (KD) of 13 μM for S2-CTP and approximately 2 to 3 μM for S2-C, S2-C-ΔR1, and S2-C-ΔR1-2 (Table S7). When NaCl was further decreased to 20 mM, KD of S2-C and S2-C-ΔR1 decreased marginally to around 0.7 μM, while that of S2-C-ΔR1-2 and S2-CTP were not significantly affected.

Figure 6

Figure 6. Investigating a possible molecular mechanism contributing to encapsulation by S2-C-derived peptides in the Cso-shell. (A) Standard free energies of binding, ΔG°, of the S2-C-derived peptides titrated against CsoS1A under various NaCl concentrations. Error bars indicate uncertainties of isotherm fitting. (B) Electrostatic potential map of a cross-sectional view of the Cso-shell lumen. Negative potential is colored red, while positive potential is represented by blue over a ±4 kT/e range. Maps were calculated using APBS. (46) (C) Electrostatic potential map of the Cso-shell exterior. (D) Calculated isoelectric points and net charge at pH 7.4 of the distinct regions on S2-C.

To probe if the shell exterior face of CsoS1A might contribute to the binding energies observed in low-ionic-strength solutions, we made use of the fact that in an intact and empty Cso-PAHA shell, only the shell exterior side of CsoS1A is able to physically interact with externally added S2-C peptides. Binding analysis at 20 and 60 mM NaCl was carried out as before, with isolated CsoS1A replaced by Cso-PAHA shells containing an equivalent concentration of CsoS1A. No binding was detected, supporting the idea that interactions of S2-C peptides with CsoS1A are confined to its luminal face. We note that the luminal face carries a distinct negative electrostatic potential (Figure 6B) due to surface exposure of multiple acidic residues. (38) Conversely, the shell exterior face of CsoS1A holds a funnel-like structure carrying a positive electrostatic potential, which has been proposed to aid in channeling of HCO3– to the central pore (Figure 6C). (44) With the exception of the R1 region, all regions of S2-C carry high isoelectric points, which is reflected in their positive net charges calculated at pH 7.4 (Figure 6D). The presence of basic regions on S2-C and the acidic character of the CsoS1A luminal face, together with the observation that moderate binding affinities (KD ∼ 1–10 μM) detected under low ionic strength were significantly weakened under a near-physiological ionic strength condition, may suggest that complementary electrostatic interactions exist between the basic regions on S2-C and the acidic CsoS1A luminal face. However, due to the unavailability of atomic-level structures of S2-C peptides in complex with CsoS1A, we are unable to provide further details on the molecular nature of binding.
The absence of detectable binding between S2-C peptides and CsoS1A under low-ionic-strength solutions in vitro appears to contradict with the observed encapsulation efficacies in vivo, in which at least two of the peptides (S2-C-ΔR1-2 and S2-CTP) could mediate cargo encapsulation into the Cso-shell at levels substantially higher than stochastic encapsulation (Table 3). We reason that while in vitro binding assessment of S2-C peptides and CsoS1A homomer may provide clues regarding the molecular characteristics of binding, it is not necessarily an accurate or sole indicator of encapsulation efficacy into the Cso-shell in vivo. We note that the length of the S2-C peptide correlates well with encapsulation efficacies (Table 3), with shorter S2-C peptides yielding higher levels of encapsulation. Full-length S2-C has been experimentally characterized to be almost entirely disordered, with only low levels of predicted local secondary structural elements. (34) Intrinsically disordered proteins tend to adopt extended conformations that occupy significantly larger volumes than globular proteins with similar masses. (45) Based on the estimated hydrodynamic radii of the S2-C peptides (Table 3), it is conceivable that the likelihood of the Cso-shell encapsulating S2-C peptides occupying smaller volumes is higher than those occupying larger volumes.
While the molecular mechanism by which S2-C peptides interact with CsoS1A is currently unascertained, it is known that in vivo, the presence of S2-CTP is sufficient for encapsulation within the Cso-shell. On the other hand, it is unknown if other repeat regions on S2-C contribute to binding under the same condition. If these regions do not contribute to binding, then their presence with S2-CTP on the same peptide sequence would only add unproductive volume onto the cargo. If there is an increase in affinity by introducing these regions, the benefit might be offset by the decreased propensity of the longer peptide to adopt a conformation that packs well into the Cso-shell before the shell is fully closed off. Hence, we propose that under intracellular conditions, sequence length of the S2-C peptide appears to be an important factor in influencing encapsulation of cargo into the Cso-shell. Given its short sequence and comparatively high GFP encapsulation efficacy, S2-CTP should be the best encapsulation peptide sequence for targeting heterologous protein cargo into the Cso-shell.

Stabilization of Enzymatic Activities Using Cso-Shells

Protein shells have been garnering attention as platforms for conferring stability on enzymes against physical insults, such as heating or freezing, or chemical insults, such as the presence of organic co-solvents or nonphysiological pH. (47,48) Enzyme confinement often reduces their conformational flexibility, which sometimes confers stability against structural changes that lead to denaturation. (49,50) Currently, homomeric protein shells are more established for hosting enzymes, attributable to their relative ease of assembly and particle size homogeneity, which improves predictability and tractability during engineering. (48,51−53) Due to their heteromeric composition, minimal BMC-derived shells represent emerging scaffolds for hosting enzymes, as these shells can provide more avenues for purposeful modifications, while their generally homogeneous particle size still confers predictability to facilitate engineering. (2,11,51,52,54) However, minimal BMC-derived shells have yet to be explored for hosting heterologous enzymes. (16,17) This encouraged us to investigate if the Cso-shell could host and stabilize enzymes. Empty Cso-shells (Cso-PAHA) were first tested for their stability against heat shock, freezing, presence of methanol co-solvent, and environments with pH from 2 to 13. The DLS spectra of shells subject to these conditions were compared to that of shells in Tris·HCl 50/350 (Tris·HCl 50 mM pH 8.0, NaCl 350 mM). Significant changes in particle size distributions and/or appearance of multiple peaks indicate protein shell disassembly. (55) Based on the conditions tested, Cso-shells were considered to be stable up to 70 °C for 15 min, 20% v/v methanol, seven consecutive freeze–thawing, and between pH 5–11 (Figure 7A–D).

Figure 7

Figure 7. Assessing the tolerance of empty Cso-shell against common denaturing factors for potential stabilization of encapsulated enzymes. (A–D) DLS spectra of empty Cso-shells tested against the conditions indicated. The baseline is vertically displaced by 0.2 for each subsequent spectrum so that all spectra can be seen in one graph. (E) Operon schematic for co-expressing Cso-shells with enzymes.

To probe the capacity of the Cso-shell for encasing enzymes of considerably different molecular sizes, an evolved pea cytosolic ascorbate peroxidase (APEX2), a 27.0 kDa monomer, and E. coli β-galactosidase (LacZ), a 466.0 kDa homotetramer, were chosen for encapsulation. (56,57) S2-CTP was fused to the C-termini of the enzymes to mediate encapsulation. Enzymes were also N-terminally tagged with the hexahistidine (His6) tag to facilitate downstream removal of unencapsulated enzymes that might co-purify with shells. (29) Cso-PAHA shells co-expressed with enzymes were constructed (Figure 7E) and purified. Immunoblotting confirmed the presence of the target enzymes in shell samples (Figure S8A,B), and the average copy numbers of enzymes per shell were estimated by Coomassie blue densitometry (Table 4). (17,29) Encapsulation of these enzymes did not appear to significantly affect Cso-shell size and morphology (Figure S8C–E).
Table 4. Quantification of the Average Copy Number of Enzymes Encapsulated Per Shell and Kinetic Constants of Encapsulated and Free Enzymesa
 average enzyme copy number per shell   
enzyme sampleT = 3T = 4kcat (s–1)KM (mM)kcat/KM × 105 (s–1·M–1)
Free APEX2N.A.N.A.460 ± 200.58 ± 0.077.93 ± 1.02
APEX2 + shell11.014.6208 ± 140.79 ± 0.122.63 ± 0.44
LacZN.A.N.A.258 ± 170 17 ± 0.0215.2 ± 2.05
LacZ + shell3.294.37200 ± 140.14 ± 0.0214.3 ± 2.27
a

For average enzyme copy number per shell, values calculated by presuming all shells are either T = 3 or 4 are provided. Kinetic measurements were performed in triplicate, and the mean values are shown with the standard error.

Encapsulation of an enzyme into a protein shell is known to alter the enzyme’s catalytic properties in some cases. To probe how encapsulation by the Cso-shell might affect the catalytic efficiencies of APEX2 and LacZ, we performed steady-state kinetics of both free and encapsulated enzymes and fitted the data to the Michaelis–Menten model to obtain the turnover number (kcat), Michaelis–Menten constant (KM), and catalytic efficiency (kcat/KM) (Table 4 and Figure S9). For encapsulated APEX2, kcat/KM decreased to approximately 30% of the free enzyme. For encapsulated LacZ, kcat/KM was not significantly different from the free enzyme. The kinetic constants, kcat and KM,, obtained for both free enzymes were in reasonable agreement with previous work, suggesting the presence of S2-CTP did not affect the activities of the free enzymes. (56,58)
Decrease in catalytic efficiencies of enzymes upon encapsulation into protein shells has been well documented and may arise from restrictions on enzyme conformational changes during substrate binding, catalysis, and product release, with the effects on kinetic parameters appearing to be enzyme-specific. (52,53,59) As an illustration, when encapsulated within the P22 capsid, kcat of an alcohol dehydrogenase decreased by approximately 85%, that of a cytochrome P450 by 30%, while that for a lipase was largely unchanged. (51−53) This is not unexpected, as different enzymes carry out catalysis via different mechanisms and conformational changes. In our study, KM of both APEX2 and LacZ were not significantly changed upon encapsulation, suggesting that substrate affinities of these enzymes were apparently not affected. On the other hand, kcat of APEX2 decreased by approximately 50% and that of LacZ by 20%. This may be attributed to loss in conformational flexibility during substrate catalysis. Impeded diffusion of substrate through the protein shell membrane into the enzyme active site has also been proposed to contribute to decrease in catalytic efficiencies. (51−53,60) A molecular dynamics study by Mahinthichaichan et al. on transport across the CsoS1A pore suggests that the pore selects for polar and anionic molecules, such as HCO3, and acts as a diffusional barrier against nonpolar molecules, such as O2 and CO2. (44) This is supported by the electrostatic potential map of the CsoS1A (Figure 6C), which depicts the pore as having a positive electrostatic potential. However, the study focused on small molecules cognate to the native α-carboxysome and the generalizability of these findings toward noncognate small molecules, as were used in this study (guaiacol and ortho-nitrophenyl-β-galactoside, ONPG), cannot be ascertained. Another molecular dynamics study on CcmK2, a β-carboxysome BMC-H protein that is structurally similar to CsoS1A, suggests that the central pores may accommodate the transit of relatively large metabolites, such as ribulose-1,5-bisphosphate, though it is likely to be less favored compared to transit of smaller metabolites. (61,62) The molecular permeability of CsoS4A, the other structural component of the Cso-shell, has not been as well investigated. Nevertheless, given that the CsoS4A pore is similar to that of CsoS1A in terms of the positive electrostatic character (Figure 6C) and size (Table S6), it would not be surprising if the permeabilities of CsoS4A and CsoS1A pores are akin to each other. Unlike in the bona fide α-carboxysome where CsoS4A is a minor component, in the Cso-shell the stoichiometry of CsoS4A is comparable to that of CsoS1A. (12) Hence, the pores of CsoS4A should be expected to significantly influence the overall permeability of the Cso-shell. In-depth investigation of the dynamics behind the permeability of the Cso-shell may provide further insights on how diffusion of small molecules across the shell barrier affects the catalytic activities of the encapsulated enzyme.
To determine possible stabilizing effects of the Cso-shell on enzymes, free enzyme and shell-encapsulated enzyme samples were challenged with the above-mentioned conditions under which empty shells were found to be stable. Enzyme activities were normalized to that of the pristine sample to determine the residual activity (Figure 8A–D). The Cso-shell conferred a moderate level of thermostability on both enzymes. Encapsulated enzymes retained approximately 90% of their activities following incubation at 40 °C for 15 min, in contrast to 40% for free enzymes. At 50 °C, encapsulated APEX2 retained about half its activity while the free enzyme was essentially inactive. However, the activity of encapsulated LacZ was only marginally higher than free enzyme at 50 °C. At 60 °C and above, all enzyme samples were inactive. The Cso-shell had a protective effect on APEX2 up to 20% v/v methanol. On the other hand, for both free and encapsulated LacZ in methanol, increases in activities were seen. It has been reported that the presence of up to 40% v/v methanol does not denature LacZ but rather enhances its activity. (63) Hence, Cso-shell was unlikely to have stabilized LacZ against methanol. For freeze–thaw stability, the Cso-shell stabilized both enzyme species up to seven consecutive cycles.

Figure 8

Figure 8. Investigating the stability of the Cso-shell and scaffolded enzymes in potentially denaturing conditions. Residual enzymes activities of free and encapsulated (+ shell) enzymes were obtained by normalizing the activity to that of the pristine sample, shown as (B) 23 °C, (C) 0% v/v methanol, (D) no freeze–thawing, and (E) pH 8. Error bars represent one standard deviation of the mean.

Encapsulated enzymes displayed higher activities within the Cso-shell at pH 10–11, but lower activities at pH 5–6. We reasoned the acidic microenvironment within the Cso-shell likely shifted the pH-activity profile of encapsulated enzymes to more alkaline conditions compared to free enzyme. The influence of anionic scaffolds on the pH-dependent activities of enzymes has been observed for synthetic maleic acid polymer scaffolds on trypsin and chymotrypsin, and more recently for the DNA polyphosphate backbone on the glucose oxidase-horseradish peroxidase (GOx-HRP) cascade. (64−66) The maleic acid polymer scaffolds increased the activities of immobilized trypsin and chymotrypsin under alkaline conditions compared to free enzymes. Similarly, the DNA scaffold emulated the acidic environment preferred by GOx and HRP, increasing the activity of the cascade relative to free enzymes in alkaline buffer. (66)
To date, the Cso-shell likely demonstrates the highest heterologous cargo loading via encapsulation peptides among minimal BMC-derived shells. The use of encapsulation peptides for such shells has been largely inefficacious and cargo often could not be detected via Coomassie blue staining, requiring more sensitive techniques such as immunoblotting or fluorescence. (16,17,30) In contrast, for the Cso-shell and S2-CTP system all three heterologous protein cargo tested (GFP, APEX2, LacZ) could be clearly identified in Coomassie-blue-stained gels (Figures S4 and S8). While orthogonal methods of cargo loading, such as structural permutation of the BMC-H subunit and installation of peptide conjugation domains into shell subunits and cargo, were developed to circumvent the hitherto low efficacy of encapsulation peptides, these peptides remain the most straightforward and widely used method for targeting protein cargo into synthetic BMCs. (17,18,24,29,37) Based on structural generalizability of BMC shell subunits, it is conceivable that for the Cso-shell, subunit engineering can be employed with encapsulation peptides for hosting of multiple cargo species in tandem, enabling more elaborate reactions to be supported by the shell. (24)

Conclusions

Click to copy section linkSection link copied!

Understanding α-carboxysome assembly is of interest as its biotechnological potential is increasingly recognized. (1,2,22) While the Cso-shell cannot fully capture the intricacies of the bona fide α-carboxysome, the establishment of a minimal shell scaffold and associated encapsulation peptides provides a platform for gaining deeper insights into the native shell. In tandem with emerging mechanistic elucidation of α-carboxysome biogenesis, the Cso-shell can be leveraged toward fully modular reconstitution and modification of the α-carboxysome, enabling applications in synthetic biology and soft matter. (34,40)
Enzyme stability often limits their utility. (47) We have shown that the Cso-shell is able to host and stabilize enzymes with markedly different molecular sizes and reaction chemistries against common denaturing factors. (47) Stabilization of APEX2 and β-galactosidase may expand their usefulness in cellular imaging and food processing applications, respectively. (56,67) Among known protein shells, a distinctive feature of the Cso-shell is its ability to stabilize encapsulated enzymes under basic environments. (48) The utility of the Cso-shell can be further explored in remediation of alkaline industrial effluent. (68) Enzymes that are currently used for wastewater treatment, such as laccases, peroxidases, and nitrilases, often suffer from diminished activities in high-pH environments. It may thus be worth exploring if these enzymes are able to benefit from encapsulation within the Cso-shell for wastewater treatment applications. (69−71) We have also identified an encapsulation peptide for the Cso-shell that mediates high levels of protein cargo encapsulation. These demonstrations advance the utility of minimal BMC-derived shells beyond its current proof-of-concept stage of encapsulating fluorescent proteins. (6,17) Characterization of the Cso-shell presented herein should inform its future employment as a bespoke nanoreactor.

Supporting Information

Click to copy section linkSection link copied!

The Supporting Information is available free of charge at https://pubs.acs.org/doi/10.1021/acs.biomac.1c00533.

  • Detailed procedures on data collection and analysis of structural models obtained from cryo-electron microscopy and X-ray crystallography; adopting and adapting an established Golden Gate-based genetic toolkit for combinatorial assembly of synthetic BMC pathways; list of constitutively active promoters from the Anderson collection; summary of recombinant combinatorial expression of Cso-shell components; multiple sequence alignment of 100 CsoS2 orthologs, focusing on the C-terminal region; and comparison of pore sizes of minimal BMC-derived shells with known atomic-scale structures and sequences of genetic constructs (PDF)

Terms & Conditions

Most electronic Supporting Information files are available without a subscription to ACS Web Editions. Such files may be downloaded by article for research use (if there is a public use license linked to the relevant article, that license may permit other uses). Permission may be obtained from ACS for other uses through requests via the RightsLink permission system: http://pubs.acs.org/page/copyright/permissions.html.

Author Information

Click to copy section linkSection link copied!

  • Corresponding Author
    • Wen Shan Yew - Department of Biochemistry, Yong Loo Lin School of Medicine, National University of Singapore (NUS), 8 Medical Drive, Singapore 117597NUS Synthetic Biology for Clinical and Technological Innovation, 28 Medical Drive, Singapore 117456Graduate School for Integrative Sciences and Engineering, NUS, Singapore 119077Synthetic Biology Translational Research Programme, Yong Loo Lin School of Medicine, National University of Singapore, 14 Medical Drive, Singapore 117599Orcidhttps://orcid.org/0000-0002-3021-0469 Email: [email protected]
  • Authors
    • Yong Quan Tan - Department of Biochemistry, Yong Loo Lin School of Medicine, National University of Singapore (NUS), 8 Medical Drive, Singapore 117597NUS Synthetic Biology for Clinical and Technological Innovation, 28 Medical Drive, Singapore 117456Graduate School for Integrative Sciences and Engineering, NUS, Singapore 119077Orcidhttps://orcid.org/0000-0002-6497-7756
    • Samson Ali - Structural Biology Research Center, Graduate School of Science, Nagoya University, Furo-cho, Chikusa-ku, Nagoya 464-8601, JapanResearch Institute for Interdisciplinary Science (RIIS), Okayama University, 3-1-1 Tsushima-naka, Kita-ku, Okayama 700-8530, Japan
    • Bo Xue - Department of Biochemistry, Yong Loo Lin School of Medicine, National University of Singapore (NUS), 8 Medical Drive, Singapore 117597NUS Synthetic Biology for Clinical and Technological Innovation, 28 Medical Drive, Singapore 117456Synthetic Biology Translational Research Programme, Yong Loo Lin School of Medicine, National University of Singapore, 14 Medical Drive, Singapore 117599
    • Wei Zhe Teo - Department of Biochemistry, Yong Loo Lin School of Medicine, National University of Singapore (NUS), 8 Medical Drive, Singapore 117597NUS Synthetic Biology for Clinical and Technological Innovation, 28 Medical Drive, Singapore 117456Synthetic Biology Translational Research Programme, Yong Loo Lin School of Medicine, National University of Singapore, 14 Medical Drive, Singapore 117599
    • Lay Hiang Ling - Department of Biochemistry, Yong Loo Lin School of Medicine, National University of Singapore (NUS), 8 Medical Drive, Singapore 117597NUS Synthetic Biology for Clinical and Technological Innovation, 28 Medical Drive, Singapore 117456Graduate School for Integrative Sciences and Engineering, NUS, Singapore 119077
    • Maybelle Kho Go - Department of Biochemistry, Yong Loo Lin School of Medicine, National University of Singapore (NUS), 8 Medical Drive, Singapore 117597NUS Synthetic Biology for Clinical and Technological Innovation, 28 Medical Drive, Singapore 117456Synthetic Biology Translational Research Programme, Yong Loo Lin School of Medicine, National University of Singapore, 14 Medical Drive, Singapore 117599
    • Hong Lv - Shanghai Engineering Research Center of Industrial Microorganisms, Shanghai 200438, People’s Republic of ChinaState Key Laboratory of Genetic Engineering, School of Life Science, Fudan University, Shanghai 200438, People’s Republic of China
    • Robert C. Robinson - Research Institute for Interdisciplinary Science (RIIS), Okayama University, 3-1-1 Tsushima-naka, Kita-ku, Okayama 700-8530, JapanSchool of Biomolecular Science and Engineering (BSE), Vidyasirimedhi Institute of Science and Technology (VISTEC), Rayong 21210, Thailand
    • Akihiro Narita - Structural Biology Research Center, Graduate School of Science, Nagoya University, Furo-cho, Chikusa-ku, Nagoya 464-8601, Japan
  • Author Contributions

    Y.Q.T., S.A., and B.X. contributed equally to this work. Y.Q.T. and W.S.Y. conceptualized the experiments. Y.Q.T., S.A., B.X., W.Z.T., and L.H.L. performed investigations. Y.Q.T., S.A., B.X., R.C.R., and A.N. conducted formal analysis on data. Y.Q.T., S.A., B.X., and W.S.Y. prepared the manuscript. M.K.G. and H.L. reviewed the manuscript. W.S.Y. supervised the project. All authors have given approval to the final version of the manuscript.

  • Funding

    This work was supported by grants from the National Research Foundation Synthetic Biology Research and Development Program (SBP-P3). A part of this work was supported by JSPS KAKENHI (18H02410) to A.N. Y.Q.T. and L.H.L. are supported by the NUS Graduate School for Integrative Sciences and Engineering scholarship.

  • Notes
    The authors declare no competing financial interest.

    A patent application (SG Non-Provisional Application No. 10202010547W) has been filed on the development and use of Cso-shells.


    The coordinates of the shells have been deposited in the Protein Data Bank (PDB) under the accession numbers 7CKB (T = 3) and 7CKC (T = 4), and Electron Microscopy Data Bank (EMDB) under the accession numbers EMD-30384 (T = 3) and EMD-30385 (T = 4). The coordinates of H. nea CsoS1D have been deposited in PDB with accession number 7DHQ.

Acknowledgments

Click to copy section linkSection link copied!

BL21(DE3)-Gold Δlac strain was a kind gift from Jeff Hasty. The authors acknowledge Dr. Lu Ting Liow for providing advice on bacterial promoters. They also thank Professors Kaoru Mitsuoka, Research Center for Ultra-High Voltage Electron Microscopy (supported by Nanotechnology Platform Program, MEXT, Japan, A-19-OS-0052), and Kenji Iwasaki, Institute for Protein Research (supported by the Cooperative Research Program of Institute for Protein Research, Osaka University), both in Osaka University for access to the cryo-electron microscopy microscopes for data optimization and collection. This research was undertaken on the MX1 beamline at the Australian Synchrotron, part of ANSTO.

Abbreviations

Click to copy section linkSection link copied!

BMC

bacterial microcompartments

BMC–P/ −T/ −H

bacterial microcompartment–pentamer/–trimer/–hexamer

DLS

dynamic light scattering

GFP

green fluorescent protein

ITC

isothermal titration calorimetry

IPTG

isopropyl β-d-1-thiogalactopyranoside

ONPG

ortho-nitrophenyl galactoside

ORF

open reading frame

PT7

T7 promoter

RuBisCO

ribulose-1,5-bisphosphate carboxylase

S2-CTP

CsoS2 C-terminal peptide

S2–N/ −M/ −C

CsoS2 N-terminal/–middle/ −C-terminal region

SII

strep-tag II

TEM

transmission electron microscopy

References

Click to copy section linkSection link copied!

This article references 71 other publications.

  1. 1
    Kerfeld, C. A.; Aussignargues, C.; Zarzycki, J.; Cai, F.; Sutter, M. Bacterial microcompartments. Nat. Rev. Microbiol. 2018, 16, 277,  DOI: 10.1038/nrmicro.2018.10
  2. 2
    Turmo, A.; Gonzalez-Esquer, C. R.; Kerfeld, C. A. Carboxysomes: metabolic modules for CO2 fixation. FEMS Microbiol. Lett. 2017, 364, fnx176  DOI: 10.1093/femsle/fnx176
  3. 3
    Shively, J. M.; Ball, F.; Brown, D. H.; Saunders, R. E. Functional Organelles in Prokaryotes: Polyhedral Inclusions (Carboxysomes) of Thiobacillus neapolitanus. Science 1973, 182, 584586,  DOI: 10.1126/science.182.4112.584
  4. 4
    Rae, B. D.; Long, B. M.; Badger, M. R.; Price, G. D. Functions, Compositions, and Evolution of the Two Types of Carboxysomes: Polyhedral Microcompartments That Facilitate CO2 Fixation in Cyanobacteria and Some Proteobacteria. Microbiol. Mol. Biol. Rev. 2013, 77, 357379,  DOI: 10.1128/MMBR.00061-12
  5. 5
    Holthuijzen, Y. A.; Kuenen, J. G.; Konings, W. N. Activity of ribulose-1,5-bisphosphate carboxylase in intact and disrupted carboxysomes of Thiobacillus neapolitanus. FEMS Microbiol. Lett. 1987, 42, 121124,  DOI: 10.1111/j.1574-6968.1987.tb02057.x
  6. 6
    Sutter, M.; Laughlin, T. G.; Sloan, N. B.; Serwas, D.; Davies, K. M.; Kerfeld, C. A. Structure of a synthetic beta-carboxysome shell. Plant Physiol. 2019, 181, 10501058,  DOI: 10.1104/pp.19.00885
  7. 7
    Sutter, M.; Greber, B.; Aussignargues, C.; Kerfeld, C. A. Assembly principles and structure of a 6.5-MDa bacterial microcompartment shell. Science 2017, 356, 12931297,  DOI: 10.1126/science.aan3289
  8. 8
    Tanaka, S.; Kerfeld, C. A.; Sawaya, M. R.; Cai, F.; Heinhorst, S.; Cannon, G. C.; Yeates, T. O. Atomic-Level Models of the Bacterial Carboxysome Shell. Science 2008, 319, 10831086,  DOI: 10.1126/science.1151458
  9. 9
    Tanaka, S.; Sawaya, M. R.; Yeates, T. O. Structure and Mechanisms of a Protein-Based Organelle in Escherichia coli. Science 2010, 327, 8184,  DOI: 10.1126/science.1179513
  10. 10
    Kalnins, G.; Cesle, E.-E.; Jansons, J.; Liepins, J.; Filimonenko, A.; Tars, K. Encapsulation mechanisms and structural studies of GRM2 bacterial microcompartment particles. Nat. Commun. 2020, 11, 388  DOI: 10.1038/s41467-019-14205-y
  11. 11
    Chowdhury, C.; Sinha, S.; Chun, S.; Yeates, T. O.; Bobik, T. A. Diverse bacterial microcompartment organelles. Microbiol. Mol. Biol. Rev. 2014, 78, 438468,  DOI: 10.1128/MMBR.00009-14
  12. 12
    Wheatley, N. M.; Gidaniyan, S. D.; Liu, Y.; Cascio, D.; Yeates, T. O. Bacterial microcompartment shells of diverse functional types possess pentameric vertex proteins. Protein Sci. 2013, 22, 660665,  DOI: 10.1002/pro.2246
  13. 13
    Hagen, A. R.; Plegaria, J. S.; Sloan, N.; Ferlez, B.; Aussignargues, C.; Burton, R.; Kerfeld, C. A. In Vitro Assembly of Diverse Bacterial Microcompartment Shell Architectures. Nano Lett. 2018, 18, 70307037,  DOI: 10.1021/acs.nanolett.8b02991
  14. 14
    Aussignargues, C.; Pandelia, M.-E.; Sutter, M.; Plegaria, J. S.; Zarzycki, J.; Turmo, A.; Huang, J.; Ducat, D. C.; Hegg, E. L.; Gibney, B. R.; Kerfeld, C. A. Structure and Function of a Bacterial Microcompartment Shell Protein Engineered to Bind a [4Fe-4S] Cluster. J. Am. Chem. Soc. 2016, 138, 52625270,  DOI: 10.1021/jacs.5b11734
  15. 15
    Lawrence, A. D.; Frank, S.; Newnham, S.; Lee, M. J.; Brown, I. R.; Xue, W.-F.; Rowe, M. L.; Mulvihill, D. P.; Prentice, M. B.; Howard, M. J.; Warren, M. J. Solution Structure of a Bacterial Microcompartment Targeting Peptide and Its Application in the Construction of an Ethanol Bioreactor. ACS Synth. Biol. 2014, 3, 454465,  DOI: 10.1021/sb4001118
  16. 16
    Cai, F.; Bernstein, S. L.; Wilson, S. C.; Kerfeld, C. A. Production and Characterization of Synthetic Carboxysome Shells with Incorporated Luminal Proteins. Plant Physiol. 2016, 170, 18681877,  DOI: 10.1104/pp.15.01822
  17. 17
    Hagen, A.; Sutter, M.; Sloan, N.; Kerfeld, C. A. Programmed loading and rapid purification of engineered bacterial microcompartment shells. Nat. Commun. 2018, 9, 2881  DOI: 10.1038/s41467-018-05162-z
  18. 18
    Fan, C.; Cheng, S.; Liu, Y.; Escobar, C. M.; Crowley, C. S.; Jefferson, R. E.; Yeates, T. O.; Bobik, T. A. Short N-terminal sequences package proteins into bacterial microcompartments. Proc. Natl. Acad. Sci. U.S.A. 2010, 107, 75097514,  DOI: 10.1073/pnas.0913199107
  19. 19
    Bonacci, W.; Teng, P. K.; Afonso, B.; Niederholtmeyer, H.; Grob, P.; Silver, P. A.; Savage, D. F. Modularity of a carbon-fixing protein organelle. Proc. Natl. Acad. Sci. U.S.A. 2012, 109, 478483,  DOI: 10.1073/pnas.1108557109
  20. 20
    Chaijarasphong, T.; Nichols, R. J.; Kortright, K. E.; Nixon, C. F.; Teng, P. K.; Oltrogge, L. M.; Savage, D. F. Programmed Ribosomal Frameshifting Mediates Expression of the Alpha-Carboxysome. J. Mol. Biol. 2016, 428, 153164,  DOI: 10.1016/j.jmb.2015.11.017
  21. 21
    Cai, F.; Dou, Z.; Bernstein, S. L.; Leverenz, R.; Williams, E. B.; Heinhorst, S.; Shively, J.; Cannon, G. C.; Kerfeld, C. A. Advances in Understanding Carboxysome Assembly in Prochlorococcus and Synechococcus Implicate CsoS2 as a Critical Component. Life 2015, 5, 11411171,  DOI: 10.3390/life5021141
  22. 22
    Li, T.; Jiang, Q.; Huang, J.; Aitchison, C. M.; Huang, F.; Yang, M.; Dykes, G. F.; He, H.-L.; Wang, Q.; Sprick, R. S.; Cooper, A. I.; Liu, L.-N. Reprogramming bacterial protein organelles as a nanoreactor for hydrogen production. Nat. Commun. 2020, 11, 5448  DOI: 10.1038/s41467-020-19280-0
  23. 23
    Kerfeld, C. A.; Melnicki, M. R. Assembly, function and evolution of cyanobacterial carboxysomes. Curr. Opin. Plant Biol. 2016, 31, 6675,  DOI: 10.1016/j.pbi.2016.03.009
  24. 24
    Plegaria, J. S.; Kerfeld, C. A. Engineering nanoreactors using bacterial microcompartment architectures. Curr. Opin. Biotechnol. 2018, 51, 17,  DOI: 10.1016/j.copbio.2017.09.005
  25. 25
    Didovyk, A.; Tonooka, T.; Tsimring, L.; Hasty, J. Rapid and Scalable Preparation of Bacterial Lysates for Cell-Free Gene Expression. ACS Synth. Biol. 2017, 6, 21982208,  DOI: 10.1021/acssynbio.7b00253
  26. 26
    Guo, Y.; Dong, J.; Zhou, T.; Auxillos, J.; Li, T.; Zhang, W.; Wang, L.; Shen, Y.; Luo, Y.; Zheng, Y.; Lin, J.; Chen, G. Q.; Wu, Q.; Cai, Y.; Dai, J. YeastFab: the design and construction of standard biological parts for metabolic engineering in Saccharomyces cerevisiae. Nucleic Acids Res. 2015, 43, e88  DOI: 10.1093/nar/gkv464
  27. 27
    Waterhouse, A. M.; Procter, J. B.; Martin, D. M. A.; Clamp, M.; Barton, G. J. Jalview Version 2─a multiple sequence alignment editor and analysis workbench. Bioinformatics 2009, 25, 11891191,  DOI: 10.1093/bioinformatics/btp033
  28. 28
    Sievers, F.; Higgins, D. G. Clustal Omega, accurate alignment of very large numbers of sequences. Methods Mol. Biol. 2014, 1079, 105116,  DOI: 10.1007/978-1-62703-646-7_6
  29. 29
    Nichols, T. M.; Kennedy, N. W.; Tullman-Ercek, D. Cargo encapsulation in bacterial microcompartments: Methods and analysis. Methods Enzymol. 2019, 617, 155186,  DOI: 10.1016/bs.mie.2018.12.009
  30. 30
    Lassila, J. K.; Bernstein, S. L.; Kinney, J. N.; Axen, S. D.; Kerfeld, C. A. Assembly of robust bacterial microcompartment shells using building blocks from an organelle of unknown function. J. Mol. Biol. 2014, 426, 22172228,  DOI: 10.1016/j.jmb.2014.02.025
  31. 31
    Anderson, J. C. Anderson Promoter Library Registry of Standard Biological Parts. http://parts.igem.org/Promoters/Catalog/Anderson (accessed 1 May, 2019).
  32. 32
    Yadav, V. G.; De Mey, M.; Lim, C. G.; Ajikumar, P. K.; Stephanopoulos, G. The future of metabolic engineering and synthetic biology: towards a systematic practice. Metab. Eng. 2012, 14, 233241,  DOI: 10.1016/j.ymben.2012.02.001
  33. 33
    Baneyx, F. Recombinant protein expression in Escherichia coli. Curr. Opin. Biotechnol. 1999, 10, 411421,  DOI: 10.1016/S0958-1669(99)00003-8
  34. 34
    Oltrogge, L. M.; Chaijarasphong, T.; Chen, A. W.; Bolin, E. R.; Marqusee, S.; Savage, D. F. Multivalent interactions between CsoS2 and Rubisco mediate α-carboxysome formation. Nat. Struct. Mol. Biol. 2020, 27, 281287,  DOI: 10.1038/s41594-020-0387-7
  35. 35
    Gonzalez, M.; Frank, E. G.; Levine, A. S.; Woodgate, R. Lon-mediated proteolysis of the Escherichia coli UmuD mutagenesis protein: in vitro degradation and identification of residues required for proteolysis. Genes Dev. 1998, 12, 38893899,  DOI: 10.1101/gad.12.24.3889
  36. 36
    Neher, S. B.; Sauer, R. T.; Baker, T. A. Distinct peptide signals in the UmuD and UmuD′ subunits of UmuD/D′ mediate tethering and substrate processing by the ClpXP protease. Proc. Natl. Acad. Sci. U.S.A. 2003, 100, 1321913224,  DOI: 10.1073/pnas.2235804100
  37. 37
    Ferlez, B.; Sutter, M.; Kerfeld, C. A. A designed bacterial microcompartment shell with tunable composition and precision cargo loading. Metab. Eng. 2019, 54, 286291,  DOI: 10.1016/j.ymben.2019.04.011
  38. 38
    Tsai, Y.; Sawaya, M. R.; Cannon, G. C.; Cai, F.; Williams, E. B.; Heinhorst, S.; Kerfeld, C. A.; Yeates, T. O. Structural Analysis of CsoS1A and the Protein Shell of the Halothiobacillus neapolitanus Carboxysome. PLoS Biol. 2007, 5, e144  DOI: 10.1371/journal.pbio.0050144
  39. 39
    Klein, M. G.; Zwart, P.; Bagby, S. C.; Cai, F.; Chisholm, S. W.; Heinhorst, S.; Cannon, G. C.; Kerfeld, C. A. Identification and structural analysis of a novel carboxysome shell protein with implications for metabolite transport. J. Mol. Biol. 2009, 392, 319333,  DOI: 10.1016/j.jmb.2009.03.056
  40. 40
    Mohajerani, F.; Sayer, E.; Neil, C.; Inlow, K.; Hagan, M. F. Mechanisms of Scaffold-Mediated Microcompartment Assembly and Size Control. ACS Nano 2021, 15, 41974212,  DOI: 10.1021/acsnano.0c05715
  41. 41
    Sinha, S.; Cheng, S.; Sung, Y. W.; McNamara, D. E.; Sawaya, M. R.; Yeates, T. O.; Bobik, T. A. Alanine scanning mutagenesis identifies an asparagine-arginine-lysine triad essential to assembly of the shell of the Pdu microcompartment. J. Mol. Biol. 2014, 426, 23282345,  DOI: 10.1016/j.jmb.2014.04.012
  42. 42
    Cai, F.; Sutter, M.; Cameron, J. C.; Stanley, D. N.; Kinney, J. N.; Kerfeld, C. A. The structure of CcmP, a tandem bacterial microcompartment domain protein from the β-carboxysome, forms a subcompartment within a microcompartment. J. Biol. Chem. 2013, 288, 1605516063,  DOI: 10.1074/jbc.M113.456897
  43. 43
    Szatmári, D.; Sárkány, P.; Kocsis, B.; Nagy, T.; Miseta, A.; Barkó, S.; Longauer, B.; Robinson, R. C.; Nyitrai, M. Intracellular ion concentrations and cation-dependent remodelling of bacterial MreB assemblies. Sci. Rep. 2020, 10, 12002  DOI: 10.1038/s41598-020-68960-w
  44. 44
    Mahinthichaichan, P.; Morris, D. M.; Wang, Y.; Jensen, G. J.; Tajkhorshid, E. Selective Permeability of Carboxysome Shell Pores to Anionic Molecules. J. Phys. Chem. B 2018, 122, 91109118,  DOI: 10.1021/acs.jpcb.8b06822
  45. 45
    Marsh, J. A.; Forman-Kay, J. D. Sequence determinants of compaction in intrinsically disordered proteins. Biophys. J. 2010, 98, 23832390,  DOI: 10.1016/j.bpj.2010.02.006
  46. 46
    Baker, N. A.; Sept, D.; Joseph, S.; Holst, M. J.; McCammon, J. A. Electrostatics of nanosystems: application to microtubules and the ribosome. Proc. Natl. Acad. Sci. U.S.A. 2001, 98, 1003710041,  DOI: 10.1073/pnas.181342398
  47. 47
    Silva, C.; Martins, M.; Jing, S.; Fu, J.; Cavaco-Paulo, A. Practical insights on enzyme stabilization. Crit. Rev. Biotechnol. 2018, 38, 335350,  DOI: 10.1080/07388551.2017.1355294
  48. 48
    Tan, Y. Q.; Xue, B.; Yew, W. S. Genetically Encodable Scaffolds for Optimizing Enzyme Function. Molecules 2021, 26, 1389  DOI: 10.3390/molecules26051389
  49. 49
    Küchler, A.; Yoshimoto, M.; Luginbühl, S.; Mavelli, F.; Walde, P. Enzymatic reactions in confined environments. Nat. Nanotechnol. 2016, 11, 409420,  DOI: 10.1038/nnano.2016.54
  50. 50
    Das, S.; Zhao, L.; Elofson, K.; Finn, M. G. Enzyme Stabilization by Virus-Like Particles. Biochemistry 2020, 59, 28702881,  DOI: 10.1021/acs.biochem.0c00435
  51. 51
    Patterson, D. P.; Schwarz, B.; El-Boubbou, K.; van der Oost, J.; Prevelige, P. E.; Douglas, T. Virus-like particle nanoreactors: programmed encapsulation of the thermostable CelB glycosidase inside the P22 capsid. Soft Matter 2012, 8, 1015810166,  DOI: 10.1039/c2sm26485d
  52. 52
    Patterson, D. P.; Prevelige, P. E.; Douglas, T. Nanoreactors by Programmed Enzyme Encapsulation Inside the Capsid of the Bacteriophage P22. ACS Nano 2012, 6, 50005009,  DOI: 10.1021/nn300545z
  53. 53
    Sánchez-Sánchez, L.; Tapia-Moreno, A.; Juarez-Moreno, K.; Patterson, D. P.; Cadena-Nava, R. D.; Douglas, T.; Vazquez-Duhalt, R. Design of a VLP-nanovehicle for CYP450 enzymatic activity delivery. J. Nanobiotechnol. 2015, 13, 66  DOI: 10.1186/s12951-015-0127-z
  54. 54
    Azuma, Y.; Zschoche, R.; Tinzl, M.; Hilvert, D. Quantitative Packaging of Active Enzymes into a Protein Cage. Angew. Chem., Int. Ed 2016, 55, 15311534,  DOI: 10.1002/anie.201508414
  55. 55
    Yu, Z.; Reid, J. C.; Yang, Y.-P. Utilizing Dynamic Light Scattering as a Process Analytical Technology for Protein Folding and Aggregation Monitoring in Vaccine Manufacturing. J. Pharm. Sci. 2013, 102, 42844290,  DOI: 10.1002/jps.23746
  56. 56
    Lam, S. S.; Martell, J. D.; Kamer, K. J.; Deerinck, T. J.; Ellisman, M. H.; Mootha, V. K.; Ting, A. Y. Directed evolution of APEX2 for electron microscopy and proximity labeling. Nat. Methods 2015, 12, 5154,  DOI: 10.1038/nmeth.3179
  57. 57
    Golan, R.; Zehavi, U.; Naim, M.; Patchornik, A.; Smirnoff, P. Inhibition of Escherichia coli beta-galactosidase by 2-nitro-1-(4,5-dimethoxy-2-nitrophenyl) ethyl, a photoreversible thiol label. Biochim. Biophys. Acta, Protein Struct. Mol. Enzymol. 1996, 1293, 238242,  DOI: 10.1016/0167-4838(95)00254-5
  58. 58
    Juers, D. H.; Hakda, S.; Matthews, B. W.; Huber, R. E. Structural Basis for the Altered Activity of Gly794 Variants of Escherichia coli β-Galactosidase. Biochemistry 2003, 42, 1350513511,  DOI: 10.1021/bi035506j
  59. 59
    Zhou, H.-X.; Rivas, G.; Minton, A. P. Macromolecular Crowding and Confinement: Biochemical, Biophysical, and Potential Physiological Consequences. Annu. Rev. Biophys. 2008, 37, 375397,  DOI: 10.1146/annurev.biophys.37.032807.125817
  60. 60
    Comellas-Aragonès, M.; Engelkamp, H.; Claessen, V. I.; Sommerdijk, N. A. J. M.; Rowan, A. E.; Christianen, P. C. M.; Maan, J. C.; Verduin, B. J. M.; Cornelissen, J. J. L. M.; Nolte, R. J. M. A virus-based single-enzyme nanoreactor. Nat. Nanotechnol. 2007, 2, 635639,  DOI: 10.1038/nnano.2007.299
  61. 61
    Faulkner, M.; Szabó, I.; Weetman, S. L.; Sicard, F.; Huber, R. G.; Bond, P. J.; Rosta, E.; Liu, L.-N. Molecular simulations unravel the molecular principles that mediate selective permeability of carboxysome shell protein. Sci. Rep. 2020, 10, 17501  DOI: 10.1038/s41598-020-74536-5
  62. 62
    Park, J.; Chun, S.; Bobik, T. A.; Houk, K. N.; Yeates, T. O. Molecular Dynamics Simulations of Selective Metabolite Transport across the Propanediol Bacterial Microcompartment Shell. J. Phys. Chem. B 2017, 121, 81498154,  DOI: 10.1021/acs.jpcb.7b07232
  63. 63
    Shifrin, S.; Hunn, G. Effect of alcohols on the enzymatic activity and subunit association of β-galactosidase. Arch. Biochem. Biophys. 1969, 130, 530535,  DOI: 10.1016/0003-9861(69)90066-6
  64. 64
    Goldstein, L.; Levin, Y.; Katchalski, E. A Water-Insoluble Polyanionic Derivative of Trypsin. II. Effect of the Polyelectrolyte Carrier on the Kinetic Behavior of the Bound Trypsin*. Biochemistry 1964, 3, 19131919,  DOI: 10.1021/bi00900a022
  65. 65
    Goldstein, L. Microenvironmental effects on enzyme catalysis. Kinetic study of polyanionic and polycationic derivatives of chymotrypsin. Biochemistry 1972, 11, 40724084,  DOI: 10.1021/bi00772a009
  66. 66
    Zhang, Y.; Tsitkov, S.; Hess, H. Proximity does not contribute to activity enhancement in the glucose oxidase–horseradish peroxidase cascade. Nat. Commun. 2016, 7, 13982  DOI: 10.1038/ncomms13982
  67. 67
    Ladero, M.; Santos, A.; García, J. L.; García-Ochoa, F. Activity over lactose and ONPG of a genetically engineered β-galactosidase from Escherichia coli in solution and immobilized: kinetic modelling. Enzyme Microb. Technol. 2001, 29, 181193,  DOI: 10.1016/S0141-0229(01)00366-0
  68. 68
    Noack, C. W.; Dzombak, D. A.; Nakles, D. V.; Hawthorne, S. B.; Heebink, L. V.; Dando, N.; Gershenzon, M.; Ghosh, R. S. Comparison of alkaline industrial wastes for aqueous mineral carbon sequestration through a parallel reactivity study. Waste Manage. 2014, 34, 18151822,  DOI: 10.1016/j.wasman.2014.03.009
  69. 69
    Arregui, L.; Ayala, M.; Gómez-Gil, X.; Gutiérrez-Soto, G.; Hernández-Luna, C. E.; Herrera de los Santos, M.; Levin, L.; Rojo-Domínguez, A.; Romero-Martínez, D.; Saparrat, M. C. N.; Trujillo-Roldán, M. A.; Valdez-Cruz, N. A. Laccases: structure, function, and potential application in water bioremediation. Microb. Cell Fact. 2019, 18, 200  DOI: 10.1186/s12934-019-1248-0
  70. 70
    Wang, M.; Abad, D.; Kickhoefer, V. A.; Rome, L. H.; Mahendra, S. Vault Nanoparticles Packaged with Enzymes as an Efficient Pollutant Biodegradation Technology. ACS Nano 2015, 9, 1093110940,  DOI: 10.1021/acsnano.5b04073
  71. 71
    Kaplan, O.; Vejvoda, V.; Plíhal, O.; Pompach, P.; Kavan, D.; Bojarová, P.; Bezouška, K.; Macková, M.; Cantarella, M.; Jirku, V.; Křen, V.; Martínková, L. Purification and characterization of a nitrilase from Aspergillus niger K10. Appl. Microbiol. Biotechnol. 2006, 73, 567575,  DOI: 10.1007/s00253-006-0503-6

Cited By

Click to copy section linkSection link copied!

This article is cited by 6 publications.

  1. Tianpei Li, Ping Chang, Weixian Chen, Zhaoyang Shi, Chunling Xue, Gregory F. Dykes, Fang Huang, Qiang Wang, Lu-Ning Liu. Nanoengineering Carboxysome Shells for Protein Cages with Programmable Cargo Targeting. ACS Nano 2024, 18 (10) , 7473-7484. https://doi.org/10.1021/acsnano.3c11559
  2. Luke M. Oltrogge, Allen W. Chen, Thawatchai Chaijarasphong, Julia B. Turnšek, David F. Savage. α-Carboxysome Size Is Controlled by the Disordered Scaffold Protein CsoS2. Biochemistry 2024, 63 (2) , 219-229. https://doi.org/10.1021/acs.biochem.3c00403
  3. Jiafeng Huang, Qiuyao Jiang, Mengru Yang, Gregory F. Dykes, Samantha L. Weetman, Wei Xin, Hai-Lun He, Lu-Ning Liu. Probing the Internal pH and Permeability of a Carboxysome Shell. Biomacromolecules 2022, 23 (10) , 4339-4348. https://doi.org/10.1021/acs.biomac.2c00781
  4. Daniel S. Trettel, Wade C. Winkler, . In Vitro Analysis of Bacterial Microcompartments and Shell Protein Superstructures by Confocal Microscopy. Microbiology Spectrum 2023, https://doi.org/10.1128/spectrum.03357-22
  5. Stephen McCarthy, Shane Gonen. Improved interface packing and design opportunities revealed by CryoEM analysis of a designed protein nanocage. Heliyon 2022, 8 (12) , e12280. https://doi.org/10.1016/j.heliyon.2022.e12280
  6. Garrett Wong, Li Rong Lim, Yong Quan Tan, Maybelle Kho Go, David J. Bell, Paul S. Freemont, Wen Shan Yew. Reconstituting the complete biosynthesis of D-lysergic acid in yeast. Nature Communications 2022, 13 (1) https://doi.org/10.1038/s41467-022-28386-6
Open PDF

Biomacromolecules

Cite this: Biomacromolecules 2021, 22, 10, 4095–4109
Click to copy citationCitation copied!
https://doi.org/10.1021/acs.biomac.1c00533
Published August 12, 2021

Copyright © 2021 The Authors. Published by American Chemical Society. This publication is licensed under

CC-BY-NC-ND 4.0 .

Article Views

2268

Altmetric

-

Citations

Learn about these metrics

Article Views are the COUNTER-compliant sum of full text article downloads since November 2008 (both PDF and HTML) across all institutions and individuals. These metrics are regularly updated to reflect usage leading up to the last few days.

Citations are the number of other articles citing this article, calculated by Crossref and updated daily. Find more information about Crossref citation counts.

The Altmetric Attention Score is a quantitative measure of the attention that a research article has received online. Clicking on the donut icon will load a page at altmetric.com with additional details about the score and the social media presence for the given article. Find more information on the Altmetric Attention Score and how the score is calculated.

  • Abstract

    Figure 1

    Figure 1. H. nea α-carboxysome operon (cso). The dotted line indicates 10 genes between cbbO and csoS1D that are unlikely to be associated with shell assembly or function. Gene lengths and intervening distances are not drawn to scale.

    Figure 2

    Figure 2. Identifying combinations that produce minimal α-carboxysome-derived shells. (A) Schematic of the combinatorial assembly process used to construct shell-expressing pathways. (B–E) Transmission electron microscopy (TEM) micrographs of combinations that produced capsid-like structures, (B) Cso-PATHA, (C) Cso-PAHA, (D) Cso-PATHC, and (E) Cso-PAHC. Scale bars (bottom right) represent 50 nm. (F, G) SDS-PAGE analysis of shell-containing samples following anion-exchange purification of (F) Cso-PATHA and (G) Cso-PAHA. The protein ladder is indicated by L in kDa.

    Figure 3

    Figure 3. Identifying S2-CR sequences for mediating cargo encapsulation within the Cso-shell. (A) CsoS2 can be split into three distinct regions based on sequence homology. Regions are not drawn to scale. (B) Schematic of the encapsulation-degradation assay. Encapsulation of GFP cargo within the shell prevents its degradation by the ClpXP protease. The blue and yellow blocks represent CsoS1A (BMC-H) and CsoS4A (BMC-P), respectively. (C) Pathway design for co-expressing UmuD (1−40)-GFP-S2-Cx (S2-Cx represents S2-C variants) with the Cso-PAHA shells. (D) Western blot detection of GFP in purified shell samples. The identities of the proteins in the ladder (L) lane are: His6-GFP-SpyTag (29.7 kDa), His6-GFP-SpyCatcher (38.2 kDa), and His6-GFP-SpyCatcher-SpyTag-GFP (67.9 kDa). Lanes 1–5 are Cso-PAHA co-expressed with the UmuD (1−40)-GFP variant indicated below the lanes. (E) DLS analysis of shells co-expressed with various cargo. The numbers to the right of the spectra correspond to the samples loaded in (D). The baseline is vertically displaced by 0.2 for each subsequent spectrum so that all peaks can be seen in one graph.

    Figure 4

    Figure 4. Atomic-scale models of the minimal α-carboxysome shells. (A) Operon schematic of Cso-PAHA + GFP-S2-CTP. (B, C) Surface representations of the shells, with the CsoS1A colored blue and CsoS4A in yellow. The blue and red patches over yellow indicate N- and C-termini of a CsoS4A monomer, and the cyan and red patches over blue indicate those of a CsoS1A monomer.

    Figure 5

    Figure 5. Close examination of the interfacial contacts in Cso-shells reinforcing notions of conserved features in BMC shell assembly. (A) CsoS4A conformations pre-shell (gray) and post-shell integration (yellow). The dashed arrow indicates movement of CsoS4A’s C-terminal loop. W80 is highlighted with dark gray (pre-integration) and orange (post-integration). (B) CsoS1A hexamers (gray) become expanded after shell incorporation (blue). Arrow indicates movement of the helices. (C) Pentamer–hexamer interface, T = 3. (D) Pentamer–hexamer interface, T = 4. (E) Hexamer–hexamer interface, T = 3. (F) Hexamer–hexamer interface, T = 4. For clarity, only half of the polar interactions left to the pseudo-2-fold axis (orange oval) at the hexamer–hexamer interface are indicated in (E) and (F). For (C)–(F), CsoS4A is colored yellow, while CsoS1A chains are colored blue and green. Dotted lines represent interatomic distances in Å.

    Figure 6

    Figure 6. Investigating a possible molecular mechanism contributing to encapsulation by S2-C-derived peptides in the Cso-shell. (A) Standard free energies of binding, ΔG°, of the S2-C-derived peptides titrated against CsoS1A under various NaCl concentrations. Error bars indicate uncertainties of isotherm fitting. (B) Electrostatic potential map of a cross-sectional view of the Cso-shell lumen. Negative potential is colored red, while positive potential is represented by blue over a ±4 kT/e range. Maps were calculated using APBS. (46) (C) Electrostatic potential map of the Cso-shell exterior. (D) Calculated isoelectric points and net charge at pH 7.4 of the distinct regions on S2-C.

    Figure 7

    Figure 7. Assessing the tolerance of empty Cso-shell against common denaturing factors for potential stabilization of encapsulated enzymes. (A–D) DLS spectra of empty Cso-shells tested against the conditions indicated. The baseline is vertically displaced by 0.2 for each subsequent spectrum so that all spectra can be seen in one graph. (E) Operon schematic for co-expressing Cso-shells with enzymes.

    Figure 8

    Figure 8. Investigating the stability of the Cso-shell and scaffolded enzymes in potentially denaturing conditions. Residual enzymes activities of free and encapsulated (+ shell) enzymes were obtained by normalizing the activity to that of the pristine sample, shown as (B) 23 °C, (C) 0% v/v methanol, (D) no freeze–thawing, and (E) pH 8. Error bars represent one standard deviation of the mean.

  • References


    This article references 71 other publications.

    1. 1
      Kerfeld, C. A.; Aussignargues, C.; Zarzycki, J.; Cai, F.; Sutter, M. Bacterial microcompartments. Nat. Rev. Microbiol. 2018, 16, 277,  DOI: 10.1038/nrmicro.2018.10
    2. 2
      Turmo, A.; Gonzalez-Esquer, C. R.; Kerfeld, C. A. Carboxysomes: metabolic modules for CO2 fixation. FEMS Microbiol. Lett. 2017, 364, fnx176  DOI: 10.1093/femsle/fnx176
    3. 3
      Shively, J. M.; Ball, F.; Brown, D. H.; Saunders, R. E. Functional Organelles in Prokaryotes: Polyhedral Inclusions (Carboxysomes) of Thiobacillus neapolitanus. Science 1973, 182, 584586,  DOI: 10.1126/science.182.4112.584
    4. 4
      Rae, B. D.; Long, B. M.; Badger, M. R.; Price, G. D. Functions, Compositions, and Evolution of the Two Types of Carboxysomes: Polyhedral Microcompartments That Facilitate CO2 Fixation in Cyanobacteria and Some Proteobacteria. Microbiol. Mol. Biol. Rev. 2013, 77, 357379,  DOI: 10.1128/MMBR.00061-12
    5. 5
      Holthuijzen, Y. A.; Kuenen, J. G.; Konings, W. N. Activity of ribulose-1,5-bisphosphate carboxylase in intact and disrupted carboxysomes of Thiobacillus neapolitanus. FEMS Microbiol. Lett. 1987, 42, 121124,  DOI: 10.1111/j.1574-6968.1987.tb02057.x
    6. 6
      Sutter, M.; Laughlin, T. G.; Sloan, N. B.; Serwas, D.; Davies, K. M.; Kerfeld, C. A. Structure of a synthetic beta-carboxysome shell. Plant Physiol. 2019, 181, 10501058,  DOI: 10.1104/pp.19.00885
    7. 7
      Sutter, M.; Greber, B.; Aussignargues, C.; Kerfeld, C. A. Assembly principles and structure of a 6.5-MDa bacterial microcompartment shell. Science 2017, 356, 12931297,  DOI: 10.1126/science.aan3289
    8. 8
      Tanaka, S.; Kerfeld, C. A.; Sawaya, M. R.; Cai, F.; Heinhorst, S.; Cannon, G. C.; Yeates, T. O. Atomic-Level Models of the Bacterial Carboxysome Shell. Science 2008, 319, 10831086,  DOI: 10.1126/science.1151458
    9. 9
      Tanaka, S.; Sawaya, M. R.; Yeates, T. O. Structure and Mechanisms of a Protein-Based Organelle in Escherichia coli. Science 2010, 327, 8184,  DOI: 10.1126/science.1179513
    10. 10
      Kalnins, G.; Cesle, E.-E.; Jansons, J.; Liepins, J.; Filimonenko, A.; Tars, K. Encapsulation mechanisms and structural studies of GRM2 bacterial microcompartment particles. Nat. Commun. 2020, 11, 388  DOI: 10.1038/s41467-019-14205-y
    11. 11
      Chowdhury, C.; Sinha, S.; Chun, S.; Yeates, T. O.; Bobik, T. A. Diverse bacterial microcompartment organelles. Microbiol. Mol. Biol. Rev. 2014, 78, 438468,  DOI: 10.1128/MMBR.00009-14
    12. 12
      Wheatley, N. M.; Gidaniyan, S. D.; Liu, Y.; Cascio, D.; Yeates, T. O. Bacterial microcompartment shells of diverse functional types possess pentameric vertex proteins. Protein Sci. 2013, 22, 660665,  DOI: 10.1002/pro.2246
    13. 13
      Hagen, A. R.; Plegaria, J. S.; Sloan, N.; Ferlez, B.; Aussignargues, C.; Burton, R.; Kerfeld, C. A. In Vitro Assembly of Diverse Bacterial Microcompartment Shell Architectures. Nano Lett. 2018, 18, 70307037,  DOI: 10.1021/acs.nanolett.8b02991
    14. 14
      Aussignargues, C.; Pandelia, M.-E.; Sutter, M.; Plegaria, J. S.; Zarzycki, J.; Turmo, A.; Huang, J.; Ducat, D. C.; Hegg, E. L.; Gibney, B. R.; Kerfeld, C. A. Structure and Function of a Bacterial Microcompartment Shell Protein Engineered to Bind a [4Fe-4S] Cluster. J. Am. Chem. Soc. 2016, 138, 52625270,  DOI: 10.1021/jacs.5b11734
    15. 15
      Lawrence, A. D.; Frank, S.; Newnham, S.; Lee, M. J.; Brown, I. R.; Xue, W.-F.; Rowe, M. L.; Mulvihill, D. P.; Prentice, M. B.; Howard, M. J.; Warren, M. J. Solution Structure of a Bacterial Microcompartment Targeting Peptide and Its Application in the Construction of an Ethanol Bioreactor. ACS Synth. Biol. 2014, 3, 454465,  DOI: 10.1021/sb4001118
    16. 16
      Cai, F.; Bernstein, S. L.; Wilson, S. C.; Kerfeld, C. A. Production and Characterization of Synthetic Carboxysome Shells with Incorporated Luminal Proteins. Plant Physiol. 2016, 170, 18681877,  DOI: 10.1104/pp.15.01822
    17. 17
      Hagen, A.; Sutter, M.; Sloan, N.; Kerfeld, C. A. Programmed loading and rapid purification of engineered bacterial microcompartment shells. Nat. Commun. 2018, 9, 2881  DOI: 10.1038/s41467-018-05162-z
    18. 18
      Fan, C.; Cheng, S.; Liu, Y.; Escobar, C. M.; Crowley, C. S.; Jefferson, R. E.; Yeates, T. O.; Bobik, T. A. Short N-terminal sequences package proteins into bacterial microcompartments. Proc. Natl. Acad. Sci. U.S.A. 2010, 107, 75097514,  DOI: 10.1073/pnas.0913199107
    19. 19
      Bonacci, W.; Teng, P. K.; Afonso, B.; Niederholtmeyer, H.; Grob, P.; Silver, P. A.; Savage, D. F. Modularity of a carbon-fixing protein organelle. Proc. Natl. Acad. Sci. U.S.A. 2012, 109, 478483,  DOI: 10.1073/pnas.1108557109
    20. 20
      Chaijarasphong, T.; Nichols, R. J.; Kortright, K. E.; Nixon, C. F.; Teng, P. K.; Oltrogge, L. M.; Savage, D. F. Programmed Ribosomal Frameshifting Mediates Expression of the Alpha-Carboxysome. J. Mol. Biol. 2016, 428, 153164,  DOI: 10.1016/j.jmb.2015.11.017
    21. 21
      Cai, F.; Dou, Z.; Bernstein, S. L.; Leverenz, R.; Williams, E. B.; Heinhorst, S.; Shively, J.; Cannon, G. C.; Kerfeld, C. A. Advances in Understanding Carboxysome Assembly in Prochlorococcus and Synechococcus Implicate CsoS2 as a Critical Component. Life 2015, 5, 11411171,  DOI: 10.3390/life5021141
    22. 22
      Li, T.; Jiang, Q.; Huang, J.; Aitchison, C. M.; Huang, F.; Yang, M.; Dykes, G. F.; He, H.-L.; Wang, Q.; Sprick, R. S.; Cooper, A. I.; Liu, L.-N. Reprogramming bacterial protein organelles as a nanoreactor for hydrogen production. Nat. Commun. 2020, 11, 5448  DOI: 10.1038/s41467-020-19280-0
    23. 23
      Kerfeld, C. A.; Melnicki, M. R. Assembly, function and evolution of cyanobacterial carboxysomes. Curr. Opin. Plant Biol. 2016, 31, 6675,  DOI: 10.1016/j.pbi.2016.03.009
    24. 24
      Plegaria, J. S.; Kerfeld, C. A. Engineering nanoreactors using bacterial microcompartment architectures. Curr. Opin. Biotechnol. 2018, 51, 17,  DOI: 10.1016/j.copbio.2017.09.005
    25. 25
      Didovyk, A.; Tonooka, T.; Tsimring, L.; Hasty, J. Rapid and Scalable Preparation of Bacterial Lysates for Cell-Free Gene Expression. ACS Synth. Biol. 2017, 6, 21982208,  DOI: 10.1021/acssynbio.7b00253
    26. 26
      Guo, Y.; Dong, J.; Zhou, T.; Auxillos, J.; Li, T.; Zhang, W.; Wang, L.; Shen, Y.; Luo, Y.; Zheng, Y.; Lin, J.; Chen, G. Q.; Wu, Q.; Cai, Y.; Dai, J. YeastFab: the design and construction of standard biological parts for metabolic engineering in Saccharomyces cerevisiae. Nucleic Acids Res. 2015, 43, e88  DOI: 10.1093/nar/gkv464
    27. 27
      Waterhouse, A. M.; Procter, J. B.; Martin, D. M. A.; Clamp, M.; Barton, G. J. Jalview Version 2─a multiple sequence alignment editor and analysis workbench. Bioinformatics 2009, 25, 11891191,  DOI: 10.1093/bioinformatics/btp033
    28. 28
      Sievers, F.; Higgins, D. G. Clustal Omega, accurate alignment of very large numbers of sequences. Methods Mol. Biol. 2014, 1079, 105116,  DOI: 10.1007/978-1-62703-646-7_6
    29. 29
      Nichols, T. M.; Kennedy, N. W.; Tullman-Ercek, D. Cargo encapsulation in bacterial microcompartments: Methods and analysis. Methods Enzymol. 2019, 617, 155186,  DOI: 10.1016/bs.mie.2018.12.009
    30. 30
      Lassila, J. K.; Bernstein, S. L.; Kinney, J. N.; Axen, S. D.; Kerfeld, C. A. Assembly of robust bacterial microcompartment shells using building blocks from an organelle of unknown function. J. Mol. Biol. 2014, 426, 22172228,  DOI: 10.1016/j.jmb.2014.02.025
    31. 31
      Anderson, J. C. Anderson Promoter Library Registry of Standard Biological Parts. http://parts.igem.org/Promoters/Catalog/Anderson (accessed 1 May, 2019).
    32. 32
      Yadav, V. G.; De Mey, M.; Lim, C. G.; Ajikumar, P. K.; Stephanopoulos, G. The future of metabolic engineering and synthetic biology: towards a systematic practice. Metab. Eng. 2012, 14, 233241,  DOI: 10.1016/j.ymben.2012.02.001
    33. 33
      Baneyx, F. Recombinant protein expression in Escherichia coli. Curr. Opin. Biotechnol. 1999, 10, 411421,  DOI: 10.1016/S0958-1669(99)00003-8
    34. 34
      Oltrogge, L. M.; Chaijarasphong, T.; Chen, A. W.; Bolin, E. R.; Marqusee, S.; Savage, D. F. Multivalent interactions between CsoS2 and Rubisco mediate α-carboxysome formation. Nat. Struct. Mol. Biol. 2020, 27, 281287,  DOI: 10.1038/s41594-020-0387-7
    35. 35
      Gonzalez, M.; Frank, E. G.; Levine, A. S.; Woodgate, R. Lon-mediated proteolysis of the Escherichia coli UmuD mutagenesis protein: in vitro degradation and identification of residues required for proteolysis. Genes Dev. 1998, 12, 38893899,  DOI: 10.1101/gad.12.24.3889
    36. 36
      Neher, S. B.; Sauer, R. T.; Baker, T. A. Distinct peptide signals in the UmuD and UmuD′ subunits of UmuD/D′ mediate tethering and substrate processing by the ClpXP protease. Proc. Natl. Acad. Sci. U.S.A. 2003, 100, 1321913224,  DOI: 10.1073/pnas.2235804100
    37. 37
      Ferlez, B.; Sutter, M.; Kerfeld, C. A. A designed bacterial microcompartment shell with tunable composition and precision cargo loading. Metab. Eng. 2019, 54, 286291,  DOI: 10.1016/j.ymben.2019.04.011
    38. 38
      Tsai, Y.; Sawaya, M. R.; Cannon, G. C.; Cai, F.; Williams, E. B.; Heinhorst, S.; Kerfeld, C. A.; Yeates, T. O. Structural Analysis of CsoS1A and the Protein Shell of the Halothiobacillus neapolitanus Carboxysome. PLoS Biol. 2007, 5, e144  DOI: 10.1371/journal.pbio.0050144
    39. 39
      Klein, M. G.; Zwart, P.; Bagby, S. C.; Cai, F.; Chisholm, S. W.; Heinhorst, S.; Cannon, G. C.; Kerfeld, C. A. Identification and structural analysis of a novel carboxysome shell protein with implications for metabolite transport. J. Mol. Biol. 2009, 392, 319333,  DOI: 10.1016/j.jmb.2009.03.056
    40. 40
      Mohajerani, F.; Sayer, E.; Neil, C.; Inlow, K.; Hagan, M. F. Mechanisms of Scaffold-Mediated Microcompartment Assembly and Size Control. ACS Nano 2021, 15, 41974212,  DOI: 10.1021/acsnano.0c05715
    41. 41
      Sinha, S.; Cheng, S.; Sung, Y. W.; McNamara, D. E.; Sawaya, M. R.; Yeates, T. O.; Bobik, T. A. Alanine scanning mutagenesis identifies an asparagine-arginine-lysine triad essential to assembly of the shell of the Pdu microcompartment. J. Mol. Biol. 2014, 426, 23282345,  DOI: 10.1016/j.jmb.2014.04.012
    42. 42
      Cai, F.; Sutter, M.; Cameron, J. C.; Stanley, D. N.; Kinney, J. N.; Kerfeld, C. A. The structure of CcmP, a tandem bacterial microcompartment domain protein from the β-carboxysome, forms a subcompartment within a microcompartment. J. Biol. Chem. 2013, 288, 1605516063,  DOI: 10.1074/jbc.M113.456897
    43. 43
      Szatmári, D.; Sárkány, P.; Kocsis, B.; Nagy, T.; Miseta, A.; Barkó, S.; Longauer, B.; Robinson, R. C.; Nyitrai, M. Intracellular ion concentrations and cation-dependent remodelling of bacterial MreB assemblies. Sci. Rep. 2020, 10, 12002  DOI: 10.1038/s41598-020-68960-w
    44. 44
      Mahinthichaichan, P.; Morris, D. M.; Wang, Y.; Jensen, G. J.; Tajkhorshid, E. Selective Permeability of Carboxysome Shell Pores to Anionic Molecules. J. Phys. Chem. B 2018, 122, 91109118,  DOI: 10.1021/acs.jpcb.8b06822
    45. 45
      Marsh, J. A.; Forman-Kay, J. D. Sequence determinants of compaction in intrinsically disordered proteins. Biophys. J. 2010, 98, 23832390,  DOI: 10.1016/j.bpj.2010.02.006
    46. 46
      Baker, N. A.; Sept, D.; Joseph, S.; Holst, M. J.; McCammon, J. A. Electrostatics of nanosystems: application to microtubules and the ribosome. Proc. Natl. Acad. Sci. U.S.A. 2001, 98, 1003710041,  DOI: 10.1073/pnas.181342398
    47. 47
      Silva, C.; Martins, M.; Jing, S.; Fu, J.; Cavaco-Paulo, A. Practical insights on enzyme stabilization. Crit. Rev. Biotechnol. 2018, 38, 335350,  DOI: 10.1080/07388551.2017.1355294
    48. 48
      Tan, Y. Q.; Xue, B.; Yew, W. S. Genetically Encodable Scaffolds for Optimizing Enzyme Function. Molecules 2021, 26, 1389  DOI: 10.3390/molecules26051389
    49. 49
      Küchler, A.; Yoshimoto, M.; Luginbühl, S.; Mavelli, F.; Walde, P. Enzymatic reactions in confined environments. Nat. Nanotechnol. 2016, 11, 409420,  DOI: 10.1038/nnano.2016.54
    50. 50
      Das, S.; Zhao, L.; Elofson, K.; Finn, M. G. Enzyme Stabilization by Virus-Like Particles. Biochemistry 2020, 59, 28702881,  DOI: 10.1021/acs.biochem.0c00435
    51. 51
      Patterson, D. P.; Schwarz, B.; El-Boubbou, K.; van der Oost, J.; Prevelige, P. E.; Douglas, T. Virus-like particle nanoreactors: programmed encapsulation of the thermostable CelB glycosidase inside the P22 capsid. Soft Matter 2012, 8, 1015810166,  DOI: 10.1039/c2sm26485d
    52. 52
      Patterson, D. P.; Prevelige, P. E.; Douglas, T. Nanoreactors by Programmed Enzyme Encapsulation Inside the Capsid of the Bacteriophage P22. ACS Nano 2012, 6, 50005009,  DOI: 10.1021/nn300545z
    53. 53
      Sánchez-Sánchez, L.; Tapia-Moreno, A.; Juarez-Moreno, K.; Patterson, D. P.; Cadena-Nava, R. D.; Douglas, T.; Vazquez-Duhalt, R. Design of a VLP-nanovehicle for CYP450 enzymatic activity delivery. J. Nanobiotechnol. 2015, 13, 66  DOI: 10.1186/s12951-015-0127-z
    54. 54
      Azuma, Y.; Zschoche, R.; Tinzl, M.; Hilvert, D. Quantitative Packaging of Active Enzymes into a Protein Cage. Angew. Chem., Int. Ed 2016, 55, 15311534,  DOI: 10.1002/anie.201508414
    55. 55
      Yu, Z.; Reid, J. C.; Yang, Y.-P. Utilizing Dynamic Light Scattering as a Process Analytical Technology for Protein Folding and Aggregation Monitoring in Vaccine Manufacturing. J. Pharm. Sci. 2013, 102, 42844290,  DOI: 10.1002/jps.23746
    56. 56
      Lam, S. S.; Martell, J. D.; Kamer, K. J.; Deerinck, T. J.; Ellisman, M. H.; Mootha, V. K.; Ting, A. Y. Directed evolution of APEX2 for electron microscopy and proximity labeling. Nat. Methods 2015, 12, 5154,  DOI: 10.1038/nmeth.3179
    57. 57
      Golan, R.; Zehavi, U.; Naim, M.; Patchornik, A.; Smirnoff, P. Inhibition of Escherichia coli beta-galactosidase by 2-nitro-1-(4,5-dimethoxy-2-nitrophenyl) ethyl, a photoreversible thiol label. Biochim. Biophys. Acta, Protein Struct. Mol. Enzymol. 1996, 1293, 238242,  DOI: 10.1016/0167-4838(95)00254-5
    58. 58
      Juers, D. H.; Hakda, S.; Matthews, B. W.; Huber, R. E. Structural Basis for the Altered Activity of Gly794 Variants of Escherichia coli β-Galactosidase. Biochemistry 2003, 42, 1350513511,  DOI: 10.1021/bi035506j
    59. 59
      Zhou, H.-X.; Rivas, G.; Minton, A. P. Macromolecular Crowding and Confinement: Biochemical, Biophysical, and Potential Physiological Consequences. Annu. Rev. Biophys. 2008, 37, 375397,  DOI: 10.1146/annurev.biophys.37.032807.125817
    60. 60
      Comellas-Aragonès, M.; Engelkamp, H.; Claessen, V. I.; Sommerdijk, N. A. J. M.; Rowan, A. E.; Christianen, P. C. M.; Maan, J. C.; Verduin, B. J. M.; Cornelissen, J. J. L. M.; Nolte, R. J. M. A virus-based single-enzyme nanoreactor. Nat. Nanotechnol. 2007, 2, 635639,  DOI: 10.1038/nnano.2007.299
    61. 61
      Faulkner, M.; Szabó, I.; Weetman, S. L.; Sicard, F.; Huber, R. G.; Bond, P. J.; Rosta, E.; Liu, L.-N. Molecular simulations unravel the molecular principles that mediate selective permeability of carboxysome shell protein. Sci. Rep. 2020, 10, 17501  DOI: 10.1038/s41598-020-74536-5
    62. 62
      Park, J.; Chun, S.; Bobik, T. A.; Houk, K. N.; Yeates, T. O. Molecular Dynamics Simulations of Selective Metabolite Transport across the Propanediol Bacterial Microcompartment Shell. J. Phys. Chem. B 2017, 121, 81498154,  DOI: 10.1021/acs.jpcb.7b07232
    63. 63
      Shifrin, S.; Hunn, G. Effect of alcohols on the enzymatic activity and subunit association of β-galactosidase. Arch. Biochem. Biophys. 1969, 130, 530535,  DOI: 10.1016/0003-9861(69)90066-6
    64. 64
      Goldstein, L.; Levin, Y.; Katchalski, E. A Water-Insoluble Polyanionic Derivative of Trypsin. II. Effect of the Polyelectrolyte Carrier on the Kinetic Behavior of the Bound Trypsin*. Biochemistry 1964, 3, 19131919,  DOI: 10.1021/bi00900a022
    65. 65
      Goldstein, L. Microenvironmental effects on enzyme catalysis. Kinetic study of polyanionic and polycationic derivatives of chymotrypsin. Biochemistry 1972, 11, 40724084,  DOI: 10.1021/bi00772a009
    66. 66
      Zhang, Y.; Tsitkov, S.; Hess, H. Proximity does not contribute to activity enhancement in the glucose oxidase–horseradish peroxidase cascade. Nat. Commun. 2016, 7, 13982  DOI: 10.1038/ncomms13982
    67. 67
      Ladero, M.; Santos, A.; García, J. L.; García-Ochoa, F. Activity over lactose and ONPG of a genetically engineered β-galactosidase from Escherichia coli in solution and immobilized: kinetic modelling. Enzyme Microb. Technol. 2001, 29, 181193,  DOI: 10.1016/S0141-0229(01)00366-0
    68. 68
      Noack, C. W.; Dzombak, D. A.; Nakles, D. V.; Hawthorne, S. B.; Heebink, L. V.; Dando, N.; Gershenzon, M.; Ghosh, R. S. Comparison of alkaline industrial wastes for aqueous mineral carbon sequestration through a parallel reactivity study. Waste Manage. 2014, 34, 18151822,  DOI: 10.1016/j.wasman.2014.03.009
    69. 69
      Arregui, L.; Ayala, M.; Gómez-Gil, X.; Gutiérrez-Soto, G.; Hernández-Luna, C. E.; Herrera de los Santos, M.; Levin, L.; Rojo-Domínguez, A.; Romero-Martínez, D.; Saparrat, M. C. N.; Trujillo-Roldán, M. A.; Valdez-Cruz, N. A. Laccases: structure, function, and potential application in water bioremediation. Microb. Cell Fact. 2019, 18, 200  DOI: 10.1186/s12934-019-1248-0
    70. 70
      Wang, M.; Abad, D.; Kickhoefer, V. A.; Rome, L. H.; Mahendra, S. Vault Nanoparticles Packaged with Enzymes as an Efficient Pollutant Biodegradation Technology. ACS Nano 2015, 9, 1093110940,  DOI: 10.1021/acsnano.5b04073
    71. 71
      Kaplan, O.; Vejvoda, V.; Plíhal, O.; Pompach, P.; Kavan, D.; Bojarová, P.; Bezouška, K.; Macková, M.; Cantarella, M.; Jirku, V.; Křen, V.; Martínková, L. Purification and characterization of a nitrilase from Aspergillus niger K10. Appl. Microbiol. Biotechnol. 2006, 73, 567575,  DOI: 10.1007/s00253-006-0503-6
  • Supporting Information

    Supporting Information


    The Supporting Information is available free of charge at https://pubs.acs.org/doi/10.1021/acs.biomac.1c00533.

    • Detailed procedures on data collection and analysis of structural models obtained from cryo-electron microscopy and X-ray crystallography; adopting and adapting an established Golden Gate-based genetic toolkit for combinatorial assembly of synthetic BMC pathways; list of constitutively active promoters from the Anderson collection; summary of recombinant combinatorial expression of Cso-shell components; multiple sequence alignment of 100 CsoS2 orthologs, focusing on the C-terminal region; and comparison of pore sizes of minimal BMC-derived shells with known atomic-scale structures and sequences of genetic constructs (PDF)


    Terms & Conditions

    Most electronic Supporting Information files are available without a subscription to ACS Web Editions. Such files may be downloaded by article for research use (if there is a public use license linked to the relevant article, that license may permit other uses). Permission may be obtained from ACS for other uses through requests via the RightsLink permission system: http://pubs.acs.org/page/copyright/permissions.html.