Skip to main content
Applied and Environmental Microbiology logoLink to Applied and Environmental Microbiology
. 2020 Nov 10;86(23):e01621-20. doi: 10.1128/AEM.01621-20

A High-Efficacy CRISPR Interference System for Gene Function Discovery in Zymomonas mobilis

Amy B Banta a,b, Amy L Enright a,b, Cheta Siletti a,*, Jason M Peters a,b,c,d,
Editor: Robert M Kellye
PMCID: PMC7657623  PMID: 32978126

Biofuels produced by microbial fermentation of plant feedstocks provide renewable and sustainable energy sources that have the potential to mitigate climate change and improve energy security. Engineered strains of the bacterium Z. mobilis can convert sugars extracted from plant feedstocks into next-generation biofuels like isobutanol; however, conversion by these strains remains inefficient due to key gaps in our knowledge about genes involved in metabolism and stress responses such as alcohol tolerance. Here, we develop CRISPRi as a tool to explore gene function in Z. mobilis. We characterize genes that are essential for growth, required to ferment sugar to ethanol, and involved in resistance to isobutanol. Our Z. mobilis CRISPRi system makes it straightforward to define gene function and can be applied to improve strain engineering and increase biofuel yields.

KEYWORDS: CRISPR-Cas9, Mismatch-CRISPRi, Mobile-CRISPRi, bioenergy, biofuel, lignocellulosic hydrolysate, essential genes, ribosomal proteins, pyruvate decarboxylase, hopanoid biosynthesis

ABSTRACT

Zymomonas mobilis is a promising biofuel producer due to its high alcohol tolerance and streamlined metabolism that efficiently converts sugar to ethanol. Z. mobilis genes are poorly characterized relative to those of model bacteria, hampering our ability to rationally engineer the genome with pathways capable of converting sugars from plant hydrolysates into valuable biofuels and bioproducts. Many of the unique properties that make Z. mobilis an attractive biofuel producer are controlled by essential genes; however, these genes cannot be manipulated using traditional genetic approaches (e.g., deletion or transposon insertion) because they are required for viability. CRISPR interference (CRISPRi) is a programmable gene knockdown system that can precisely control the timing and extent of gene repression, thus enabling targeting of essential genes. Here, we establish a stable, high-efficacy CRISPRi system in Z. mobilis that is capable of perturbing all genes—including essential genes. We show that Z. mobilis CRISPRi causes either strong knockdowns (>100-fold) using single guide RNA (sgRNA) spacers that perfectly match target genes or partial knockdowns using spacers with mismatches. We demonstrate the efficacy of Z. mobilis CRISPRi by targeting essential genes that are universally conserved in bacteria, are key to the efficient metabolism of Z. mobilis, or underlie alcohol tolerance. Our Z. mobilis CRISPRi system will enable comprehensive gene function discovery, opening a path to rational design of biofuel production strains with improved yields.

IMPORTANCE Biofuels produced by microbial fermentation of plant feedstocks provide renewable and sustainable energy sources that have the potential to mitigate climate change and improve energy security. Engineered strains of the bacterium Z. mobilis can convert sugars extracted from plant feedstocks into next-generation biofuels like isobutanol; however, conversion by these strains remains inefficient due to key gaps in our knowledge about genes involved in metabolism and stress responses such as alcohol tolerance. Here, we develop CRISPRi as a tool to explore gene function in Z. mobilis. We characterize genes that are essential for growth, required to ferment sugar to ethanol, and involved in resistance to isobutanol. Our Z. mobilis CRISPRi system makes it straightforward to define gene function and can be applied to improve strain engineering and increase biofuel yields.

INTRODUCTION

Zymomonas mobilis is a Gram-negative alphaproteobacterium with superlative properties for biofuel production (15) but poorly characterized gene functions (6). Z. mobilis is an efficient, natural ethanologen (7, 8) capable of fermenting glucose to ethanol at 97% of the theoretical yield (7, 9) with little energy spent on biomass (8, 10). Furthermore, Z. mobilis is highly resistant to ethanol, up to 16% (vol/vol) (8). However, our ability to engineer strains of Z. mobilis that produce high yields of advanced biofuels, such as isobutanol (IBA), has been hindered by the lack of functional information for two key gene sets: metabolic and stress response/resistance genes. Genetic analysis to identify and characterize metabolic and stress response genes could allow us to engineer strains with increased flux toward IBA and away from ethanol (3, 11), as well as strains that are resistant to hydrolysate inhibitors like acetic acid and various phenolic compounds (4).

Z. mobilis has a minimalistic metabolism with little functional redundancy (1214). Z. mobilis converts sugars to pyruvate via the Entner-Doudoroff (ED) pathway, rather than the more commonly used but less thermodynamically favorable Embden-Meyerhof-Parnas (EMP) pathway (1518). Metabolic models based on Z. mobilis ZM4 genome sequences (1214, 18) revealed that central metabolic pathways like EMP glycolysis and the tricarboxylic acid (TCA) cycle are missing key enzymes (e.g., 6-phosphofructokinase and 2-oxoglutarate dehydrogenase, respectively), further limiting its metabolic plasticity. Because of its streamlined metabolism, many metabolic genes are predicted to be essential for growth in Z. mobilis.

Another defining feature of Z. mobilis physiology is the production of large quantities of hopanoids (19), i.e., triterpenoid lipids that provide resistance to environmental stresses in bacteria (2022). Hopanoids are thought to act by altering membrane fluidity and permeability, analogous to the action of cholesterol—also a triterpenoid lipid—on eukaryotic membranes (2325). Although other bacteria make hopanoids, Z. mobilis produces them in higher quantities, with the number of hopanoids nearly matching that of phospholipids in the cell during peak production conditions (19). Hopanoids are thought to be essential to Z. mobilis, as chemical inhibition of enzymes involved in hopanoid precursor biosynthesis inhibits growth (26, 27) and transposon insertions into hopanoid biosynthesis genes result in cells with both a wild-type and transposon mutant allele (i.e., strains with both hpn+ and hpn::Tn alleles) (28, 29). Consistent with having an essential role in Z. mobilis physiology, hopanoid production is correlated with the ethanol content of the growth medium (30) and mutations in hopanoid biosynthesis genes increase sensitivity to ethanol (28). Whether hopanoids are required for resistance to additional stresses, such as hydrolysate toxins or alcohols other than ethanol, is unknown.

CRISPR interference (CRISPRi)—the use of programmable guide RNAs and Cas proteins (31, 32) or complexes (33, 34) lacking nuclease activity to repress transcription of target genes—is capable of probing the functions of essential genes. This is because CRISPRi knockdowns are inducible and titratable (31, 32, 35), separating the steps of strain construction and gene phenotyping. CRISPRi has been used to phenotype essential genes in multiple bacterial species (36) and define chemical-gene interactions (35, 37), cell morphology phenotypes (35, 38, 39), host genes involved in phage life cycles (40), novel gene functions (35, 38), and essential gene network architecture (35), among other purposes. To facilitate essential gene phenotyping by CRISPRi in diverse bacterial species, we recently developed “Mobile-CRISPRi”—a suite of modular CRISPRi vectors based on the extensively studied type II-A CRISPR system from Streptococcus pyogenes (i.e., Streptococcus pyogenes catalytically dead Cas9 [Spy dCas9]) (Fig. 1A) that are transferred by mating and site specifically integrate into the genomes of recipient bacteria (41, 42). We demonstrated integration and knockdown in species ranging from Gram-negative Gammaproteobacteria to Gram-positive Firmicutes (41), but we did not test species from Alphaproteobacteria. Z. mobilis contains a native type I to F CRISPR system that has been co-opted for efficient genome editing, as well as CRISPRi in a Cas3 nuclease-deficient background (34); however, CRISPRi using the native system showed low knockdown efficacy (4- to 5-fold maximum) and was not inducible as constructed, limiting its usefulness in targeting essential genes. We reasoned that importing the heterologous Spy dCas9 system would result in stronger knockdowns, as is seen in other species (31, 35, 41).

FIG 1.

FIG 1

Mobile-CRISPRi system for transcriptional repression optimized for Zymomonas mobilis. (A) Modular Z. mobilis CRISPRi system encodes dCas9, sgRNA, and antibiotic resistance cassettes on a Tn7 transposon. The promoter (P1) and ribosome binding site (rbs) for dCas9 and the promoter (PC) for the sgRNA have been optimized for Z. mobilis. DNA encoding the 20-nt variable region of the sgRNA can be cloned (individually or libraries) in between the BsaI sites. (B) CRISPRi-expressing strains are constructed by triparental mating of E. coli donor strains (one harboring the Mobile-CRISPRi plasmid and another harboring a plasmid expressing the Tn7 transposase) with Z. mobilis. The CRISPRi expression cassette will be stably incorporated onto the Z. mobilis chromosome at the Tn7 att site located downstream from glmS. (C) Optimization of sgRNA expression. Six promoter sequences (A to F) based on either lacUV5 or a synthetic promoter were incorporated into the CRISPRi system. Alignment is to the E. coli σ70 consensus promoter, with the −10 and −35 core promoter elements underlined and shown in boldface, the lac operator locations highlighted in green or cyan, and the UP element highlighted in orange. (D) Comparison of Z. mobilis Mobile-CRISPRi sgRNA promoter variants. A GFP expression cassette was cloned into the PmeI site, and an sgRNA targeting GFP (or a nontargeting control) was cloned into the BsaI sites. Cultures were diluted 1:1,000 and incubated in medium with 0 or 1 mM IPTG for ∼10 doublings prior to measurement of GFP expression. Expression was normalized to the results for a non GFP-expressing strain. Standard deviations of the results from 4 biological replicates are shown. (E) Expression of Z. mobilis CRISPRi system is inducible over a range of IPTG concentrations. Standard deviations are shown.

Here, we establish a stable and efficacious CRISPRi system for Z. mobilis based on Spy dCas9. We demonstrate strong (>100-fold) or partial knockdown of gene expression by using single guide RNA (sgRNA) spacers that are complementary or mismatched to target genes, respectively. We use Z. mobilis CRISPRi to demonstrate the essentiality of genes involved in metabolism and hopanoid biosynthesis. Furthermore, we show that reduced expression of specific hopanoid biosynthesis genes leads to IBA sensitivity. Our Z. mobilis CRISPRi system will enable rapid characterization of gene function, accelerating rational engineering of the genome for advanced biofuel production.

RESULTS

Optimization of CRISPRi for Z. mobilis.

To establish Spy dCas9-based CRISPRi in Z. mobilis, we first attempted to deliver a previously described Tn7-based Mobile-CRISPRi test construct containing the gene encoding monomeric red fluorescent protein (mRFP) and an sgRNA targeting mRFP (41) to wild-type Z. mobilis strain ZM4 via conjugation (Fig. 1A and B). We failed to obtain transconjugants with the wild type but succeeded at integrating Mobile-CRISPRi into the genome of a restriction-deficient derivative strain (Piyush B. Lal and Patricia J. Kiley, unpublished data), consistent with Mobile-CRISPRi vectors containing multiple predicted recognition sites for Z. mobilis restriction enzymes (43). Therefore, we used the restriction-deficient strain in all subsequent experiments (Table 1). Mobile-CRISPRi inserted into the Z. mobilis genome downstream from glmS, as expected (Fig. S1 in the supplemental material), and was stable over 50 generations of growth in rich medium without selection. We next used a fluorometer to measure CRISPRi knockdown of mRFP at saturating concentrations of inducer (1 mM IPTG [isopropyl-β-d-thiogalactopyranoside]), finding poor knockdown (2.4-fold) (Fig. S2), although our measurements were complicated by the weak fluorescence of mRFP in Z. mobilis.

TABLE 1.

Strains

Strain Descriptiona Reference or source
sJMP006 Escherichia coli (Keio WT strain) (BW25113) F λ lacIq rrnBT14 ΔlacZWJ16 hsdR514 ΔaraBADAH33 ΔrhaBADLD78 62
sJMP032 Escherichia coli DH10B (cloning strain) F Δ(ara-leu)7697[Δ(rapA′-cra′)] Δ(lac)X74[Δ(′yahH-mhpE)] duplication(514341–627601)[nmpC-gltI] galK16 galE15 e14(icdWT mcrA) ϕ80dlacZΔM15 recA1 relA1 endA1 Tn10.10 nupG rpsL150(Strr) rph+ spoT1 Δ(mrr-hsdRMS-mcrBC) λ Missense(dnaA glmS glyQ lpxK mreC murA) Nonsense(chiA gatZ fhuA? yigA ygcG) Frameshift(flhC mglA fruB) Invitrogen
sJMP146 Escherichia coli (pir+ cloning strain) (BW25141) Δ(araD-araB)567 ΔlacZ4787(::rrnB-3) Δ(phoB-phoR)580 λ galU95 ΔuidA3::pir+ recA1 endA9(Δ-ins)::FRT rph-1 Δ(rhaD-rhaB)568 hsdR514 65
sJMP412 Zymomonas mobilis ZM4 ΔhsdSC (ZMO1933) Δmrr (ZMO0028) ΔhsdSR (pZM32_028) Δcas3 (ZMO0681) (strain PK15436) Lal and Kiley, unpublished data
sJMP424 Escherichia coli (pir+ dap-negative mating strain) (strain WM6026) lacIq rrnB3 DElacZ4787 hsdR514 DE(araBAD)567 DE(rhaBAD)568 rph-1 att-lambda::pAE12-del (oriR6K/cat::FRT5) Δ4229(dapA)::FRT(DAP) Δ(endA)::FRT uidAMluI)::pir(wt) attHK::pJK1006::Δ1/2(ΔoriR6K-cat::FRT5 ΔtrfA::FRT) 66
sJMP2032 sJMP412 with CRISPRi system from pJMP196 in Tn7 att, Kanr (MCi-RFP-rfp) This study
sJMP2035 sJMP412 with CRISPRi system from pJMP197 in Tn7 att, Kanr (MCi-RFP-BsaI) This study
sJMP2065 sJMP412 with CRISPRi system from pJMP2044 in Tn7 att, Cmr (MCi-noGFP) This study
sJMP2069 sJMP412 with CRISPRi system from pJMP2046 in Tn7 att, Cmr (MCi-vA-GFP-BsaI) This study
sJMP2073 sJMP412 with CRISPRi system from pJMP2048 in Tn7 att, Cmr (MCi-vA-GFP-gfp) This study
sJMP2118 sJMP412 with CRISPRi system from pJMP2093 in Tn7 att, Cmr (MCi-vB-GFP-BsaI) This study
sJMP2122 sJMP412 with CRISPRi system from pJMP2095 in Tn7 att, Cmr (MCi-vB-GFP-gfp) This study
sJMP2340 sJMP2065 with plasmid pSRK-kan (pJMP2316) This study
sJMP2341 sJMP2065 with plasmid pSRK-kan-sgRNA (pJMP2317) This study
sJMP2342 sJMP2065 with plasmid pSRK-kan-dCas9 (pJMP2319) This study
sJMP2343 sJMP2069 with plasmid pSRK-kan (pJMP2316) This study
sJMP2344 sJMP2069 with plasmid pSRK-kan-sgRNA (pJMP2317) This study
sJMP2345 sJMP2069 with plasmid pSRK-kan-dCas9 (pJMP2319) This study
sJMP2346 sJMP2073 with plasmid pSRK-kan (pJMP2316) This study
sJMP2347 sJMP2073 with plasmid pSRK-kan-sgRNA (pJMP2317) This study
sJMP2348 sJMP2073 with plasmid pSRK-kan-dCas9 (pJMP2319) This study
sJMP2349 sJMP2118 with plasmid pSRK-kan (pJMP2316) This study
sJMP2350 sJMP2118 with plasmid pSRK-kan-sgRNA (pJMP2317) This study
sJMP2351 sJMP2118 with plasmid pSRK-kan-dCas9 (pJMP2319) This study
sJMP2352 sJMP2122 with plasmid pSRK-kan (pJMP2316) This study
sJMP2353 sJMP2122 with plasmid pSRK-kan-sgRNA (pJMP2317) This study
sJMP2354 sJMP2122 with plasmid pSRK-kan-dCas9 (pJMP2319) This study
sJMP2430 sJMP412 with CRISPRi system from pJMP2367 in Tn7 att, Cmr (MCi-vC-GFP-BsaI) This study
sJMP2433 sJMP412 with CRISPRi system from pJMP2369 in Tn7 att, Cmr (MCi-vD-GFP-BsaI) This study
sJMP2436 sJMP412 with CRISPRi system from pJMP2371 in Tn7 att, Cmr (MCi-vE-GFP-BsaI) This study
sJMP2439 sJMP412 with CRISPRi system from pJMP2373 in Tn7 att, Cmr (MCi-vF-GFP-BsaI) This study
sJMP2442 sJMP412 with CRISPRi system from pJMP2375 in Tn7 att, Cmr (MCi-vC-GFP-gfp) This study
sJMP2445 sJMP412 with CRISPRi system from pJMP2377 in Tn7 att, Cmr (MCi-vD-GFP-gfp) This study
sJMP2447 sJMP412 with CRISPRi system from pJMP2379 in Tn7 att, Cmr (MCi-vE-GFP-gfp) This study
sJMP2451 sJMP412 with CRISPRi system from pJMP2381 in Tn7 att, Cmr (MCi-vF-GFP-gfp) This study
sJMP2454 sJMP412 with CRISPRi system from pJMP2391 in Tn7 att, Cmr (hpnC sgRNA) This study
sJMP2456 sJMP412 with CRISPRi system from pJMP2393 in Tn7 att, Cmr (hpnC sgRNA) This study
sJMP2458 sJMP412 with CRISPRi system from pJMP2395 in Tn7 att, Cmr (hpnF sgRNA) This study
sJMP2460 sJMP412 with CRISPRi system from pJMP2397 in Tn7 att, Cmr (hpnF sgRNA) This study
sJMP2462 sJMP412 with CRISPRi system from pJMP2399 in Tn7 att, Cmr (hpnH sgRNA) This study
sJMP2463 sJMP412 with CRISPRi system from pJMP2401 in Tn7 att, Cmr (hpnH sgRNA) This study
sJMP2465 sJMP412 with CRISPRi system from pJMP2403 in Tn7 att, Cmr (hpnI sgRNA) This study
sJMP2467 sJMP412 with CRISPRi system from pJMP2405 in Tn7 att, Cmr (hpnI sgRNA) This study
sJMP2469 sJMP412 with CRISPRi system from pJMP2407 in Tn7 att, Cmr (shc2 sgRNA) This study
sJMP2471 sJMP412 with CRISPRi system from pJMP2409 in Tn7 att, Cmr (shc2 sgRNA) This study
sJMP2477 sJMP412 with CRISPRi system from pJMP2415 in Tn7 att, Cmr (rfp sgRNA) This study
sJMP2543 sJMP412 with CRISPRi system from pJMP2367 in Tn7 att, Cmr (MCi-vC-GFP-BsaI) This study
sJMP2544 sJMP412 with CRISPRi system from pJMP2375 in Tn7 att, Cmr (MCi-vC-GFP-gfp) This study
sJMP2545 sJMP412 with CRISPRi system from pJMP2409 in Tn7 att, Cmr (MCi-vC-GFP-gmc1) This study
sJMP2546 sJMP412 with CRISPRi system from pJMP2411 in Tn7 att, Cmr (MCi-vC-GFP-gmc2) This study
sJMP2547 sJMP412 with CRISPRi system from pJMP2413 in Tn7 att, Cmr (MCi-vC-GFP-gmc3) This study
sJMP2548 sJMP412 with CRISPRi system from pJMP2415 in Tn7 att, Cmr (MCi-vC-GFP-gmc4) This study
sJMP2549 sJMP412 with CRISPRi system from pJMP2417 in Tn7 att, Cmr (MCi-vC-GFP-gmc5) This study
sJMP2550 sJMP412 with CRISPRi system from pJMP2419 in Tn7 att, Cmr (MCi-vC-GFP-gmc6) This study
sJMP2551 sJMP412 with CRISPRi system from pJMP2421 in Tn7 att, Cmr (MCi-vC-GFP-gmc7) This study
sJMP2552 sJMP412 with CRISPRi system from pJMP2423 in Tn7 att, Cmr (MCi-vC-GFP-gmc8) This study
sJMP2553 sJMP412 with CRISPRi system from pJMP2425 in Tn7 att, Cmr (MCi-vC-GFP-gmc9) This study
sJMP2554 sJMP412 with CRISPRi system from pJMP2480 in Tn7 att, Cmr (MCi-vC-noGFP-BsaI) This study
sJMP2555 sJMP006 with CRISPRi system from pJMP2367 in Tn7 att, Cmr This study
sJMP2556 sJMP006 with CRISPRi system from pJMP2375 in Tn7 att, Cmr This study
sJMP2557 sJMP006 with CRISPRi system from pJMP2409 in Tn7 att, Cmr This study
sJMP2558 sJMP006 with CRISPRi system from pJMP2411 in Tn7 att, Cmr This study
sJMP2559 sJMP006 with CRISPRi system from pJMP2413 in Tn7 att, Cmr This study
sJMP2560 sJMP006 with CRISPRi system from pJMP2415 in Tn7 att, Cmr This study
sJMP2561 sJMP006 with CRISPRi system from pJMP2417 in Tn7 att, Cmr This study
sJMP2562 sJMP006 with CRISPRi system from pJMP2419 in Tn7 att, Cmr This study
sJMP2563 sJMP006 with CRISPRi system from pJMP2421 in Tn7 att, Cmr This study
sJMP2564 sJMP006 with CRISPRi system from pJMP2423 in Tn7 att, Cmr This study
sJMP2565 sJMP006 with CRISPRi system from pJMP2425 in Tn7 att, Cmr This study
sJMP2566 sJMP006 with CRISPRi system from pJMP2480 in Tn7 att, Cmr This study
sJMP2605 sJMP412 with CRISPRi system from pJMP2597 in Tn7 att, Cmr This study
sJMP2606 sJMP412 with CRISPRi system from pJMP2598 in Tn7 att, Cmr This study
sJMP2607 sJMP412 with CRISPRi system from pJMP2599 in Tn7 att, Cmr This study
sJMP2608 sJMP412 with CRISPRi system from pJMP2600 in Tn7 att, Cmr This study
a

WT, wild type; Cmr, chloramphenicol resistance cassette; Ampr, ampicillin resistance cassette; Kanr, kanamycin resistance cassette; Specr, spectinomycin resistance cassette; MCi, Mobile-CRISPRi; RFP, red fluorescent protein; vA to vF, sgRNA promoter variants; sfGFP, superfolder GFP.

To optimize CRISPRi function and improve knockdown detection, we took advantage of the modularity of Mobile-CRISPRi (Fig. 1A) to swap in biological parts that have been confirmed to function in Z. mobilis (44). We replaced mRFP with the gene encoding superfolder green fluorescent protein (GFP) (sfGFP) (45), expressed dcas9 from a T7A1-derived promoter with a strong Z. mobilis ribosome binding site, and expressed an sgRNA targeting sfGFP from the lacUV5 promoter (i.e., promoter A) (Fig. 1C). This CRISPRi system showed strong knockdown (28-fold) of sfGFP at a saturating inducer concentration, but also considerable leakiness (9-fold) without inducer (Fig. 1D, promoter A). Inserting a symmetric lac operator site (46) into the lacUV5 promoter spacer (Fig. 1C, promoter B) resulted in no detectable leakiness but only modest knockdown (9-fold) (Fig. 1D), suggesting that the concentration of either dCas9 or the sgRNA was limiting for knockdown. To determine the limiting factor, we expressed either the sfGFP sgRNA or dcas9 from a multicopy plasmid in the context of CRISPRi with promoter B and found that sgRNA expression was primarily limiting knockdown (Fig. S3). Because lacUV5 has the highest confirmed activity of any promoter measured in Z. mobilis (44) and because the Z. mobilis promoter consensus sequence (47) was unknown when we began our studies, we built LacI-regulated synthetic promoters based on Escherichia coli σ70 consensus elements (48, 49) that we reasoned could increase sgRNA expression to a higher level than lacUV5 (Fig. 1C, promoters C to F). All four synthetic promoters improved the knockdown properties of Z. mobilis CRISPRi, but promoter C, which features consensus UP and −10 elements with a near-consensus −35 and an ideal spacer length (Fig. 1C), provided the best combination of strong knockdown (125-fold) and negligible leakiness (∼10 to 15%) (Fig. 1D). Using CRISPRi with promoter C, we found that intermediate inducer concentrations enabled titration of knockdown activity (Fig. 1E). We conclude that Mobile-CRISPRi optimized for Z. mobilis is efficacious, inducible, and titratable.

Mismatch-CRISPRi enables knockdown gradients in Z. mobilis.

The relationship between fitness and gene expression varies by gene and is generally unknown (50, 51). This relationship is especially important to consider for essential genes, which have a fitness of zero at full knockdown but a large range of possible fitness values at intermediate levels of knockdown, depending on the function of the gene product. Excessive knockdown of essential genes results in strains that grow poorly and are difficult to phenotype. The Gross and Weissman laboratories have shown that systematically introducing mismatches between sgRNA spacers and target genes can generate knockdown gradients suitable for studying essential gene function (5053); we call this strategy “Mismatch-CRISPRi.” Mismatch-CRISPRi functions in mammalian cells (51) and diverse model bacteria (i.e., E. coli and Bacillus subtilis [50]), although the behavior of mismatched guides is not identical in mammalian and bacterial systems (50). Hawkins and Silvis et al. took a machine learning approach to characterize the behavior of mismatched guides targeting sfGFP (∼1,500) in both E. coli and B. subtilis, finding that ∼50% of the mismatched-guide activity could be predicted by taking into account the mismatch position, changes in the free energy of sgRNA-DNA pairing, and %GC content; these predictions were validated by targeting essential genes with mismatched guides in both organisms (50). To test whether mismatched sgRNAs behave similarly in Alphaproteobacteria, we cloned a subset of the mismatched guides used by Hawkins and Silvis et al. to target sfGFP (50) into our Z. mobilis CRISPRi system with sgRNA promoter C (Fig. 2A and B). Using these mismatched guides, we were able to generate a knockdown gradient of sfGFP that spanned nearly 2 orders of magnitude and contained multiple sgRNAs that caused intermediate levels of knockdown at a saturating inducer concentration (Fig. 2C). Furthermore, we introduced our Z. mobilis Mismatch-CRISPRi vectors into E. coli, permitting a direct comparison of sfGFP knockdown in the two divergent species (Fig. S4). Consistent with a previous comparison between E. coli and B. subtilis (50), we found excellent agreement between sfGFP knockdown gradients in E. coli and Z. mobilis (R2 = 0.8). This demonstrates the broad utility of Mismatch-CRISPRi to predictably generate partial knockdowns in diverse bacteria and suggests that Z. mobilis CRISPRi may function well in multiple species.

FIG 2.

FIG 2

Variable levels of repression using mismatched sgRNAs in the Z. mobilis Mobile-CRISPRi system. (A) Location of sgRNA targets gmc1 to gmc9 (GFP Mismatch-CRISPRi guide RNAs) on the GFP gene. Scale bar indicates nucleotides. (B) Sequences of the GFP-targeting sgRNAs, with mismatches indicated in lowercase red. Protospacer adjacent motif (PAM)-proximal seed sequence is indicated. (C) Knockdown of GFP expression in Z. mobilis CRISPRi expression strains with mismatched sgRNAs. “Control” indicates a nontargeting sgRNA. Standard deviations are shown. IPTG (1 mM) was used for induction.

Z. mobilis CRISPRi targets essential genes.

To examine the efficacy of Z. mobilis CRISPRi in characterizing essential gene function, we first targeted rplL (ZMO0728)—an essential gene encoding the universally conserved ribosomal protein L12 (54)—as a positive control. We found a reduction greater than 6 orders of magnitude in plating efficiency for strains expressing an rplL sgRNA versus a control strain expressing a nontargeting sgRNA at saturating inducer (Fig. 3A), indicating substantial loss of cell viability and relatively low levels of suppressor mutations that inactivate the CRISPRi system. Based on these results, we conclude that Z. mobilis CRISPRi is effective at assessing gene essentiality and allowing observation of essential gene knockdown phenotypes.

FIG 3.

FIG 3

CRISPRi knockdown of endogenous essential genes. Z. mobilis strains with CRISPRi cassettes encoding sgRNAs targeting essential genes rplL (A) and pdc (B and C) were serially diluted 1:10 (10−1 through 10−8) and spotted on agar plates with either 0 or 1 mM IPTG. “C” indicates a nontargeting sgRNA (control).

Pyruvate decarboxylase, encoded by the pdc gene (ZMO1360), is a key metabolic enzyme in Z. mobilis that converts pyruvate into acetaldehyde—the penultimate step in ethanol production (55). Despite the important role of pdc in fermentation of sugars to ethanol, the Z. mobilis literature is conflicted about whether pdc is essential (5, 56) or dispensable (57, 58) under aerobic conditions. To determine the essentiality of pdc, we used Z. mobilis CRISPRi with promoter C and an sgRNA targeting the 5′ end of the pdc coding sequence. We found a loss in plating efficiency greater than 6 orders of magnitude for the pdc knockdown strain at a saturating inducer concentration (Fig. 3B); this result was indistinguishable from the loss of fitness observed when we targeted rplL, suggesting that pdc is essential for aerobic growth. To confirm that our result was not due to off-target effects of CRISPRi, we tested a second, nonoverlapping sgRNA targeting pdc, finding the same results (Fig. 3C). We conclude that pdc is essential for aerobic growth of Z. mobilis.

Essentiality and IBA sensitivity phenotypes of hopanoid biosynthesis genes.

Genes encoding hopanoid biosynthesis enzymes (i.e., hpn and shc genes) (Fig. 4A and Fig. S5) are thought to be essential in Z. mobilis, based on growth cessation caused by small-molecule inhibitors of squalene-hopene cyclase (26) and the observation that strains with transposon insertions in hpn genes always also contain a wild-type copy of the gene (28). To further probe the essentiality of hopanoids, we targeted hpn and shc genes using Z. mobilis CRISPRi. Because CRISPRi blocks transcription of downstream genes in an operon (i.e., polarity), we chose to target the first hpn or shc gene present in each operon (Fig. 4A, orange genes). We found considerable defects in plating efficiency for strains with sgRNAs targeting hpnC (ZMO0869), hpnH (ZMO0874), and hpnI (ZMO0972), consistent with a requirement of hopanoid synthesis for growth (Fig. 4B). In contrast, targeting the hpnF (also known as shc1 [ZMO0872]) and shc2 (ZMO1548) genes that both encode squalene-hopene cyclase had no effect on plating efficiency, suggesting that they are functionally redundant under the conditions tested.

FIG 4.

FIG 4

CRISPRi knockdown of hopanoid lipid synthesis-related genes. (A) Z. mobilis strains were constructed with CRISPRi cassettes encoding sgRNAs targeting genes in hopanoid synthesis operons (and a nontargeting control). The number of the ZMO locus tag is in parentheses after the gene name. Targeted genes (hpnC, hpnF [shc1], hpnH, hpnI, and shc2) are shown in orange. Target positions of sgRNAs are shown in black under the genes. (B) Strains were serially diluted 1:10 (10−2 through 10−8) and spotted on agar plates with 0, 0.01, 0.1, or 1 mM IPTG. “C” indicates a nontargeting sgRNA (control). (C) Strains were diluted 1:1,000 and grown for ∼10 doublings in liquid culture, aerobically, in the presence of 1.25% isobutanol and 0 or 0.1 mM IPTG prior to measurement of cell density (OD600). Growth measurements were normalized to the growth of a strain expressing a nontargeting sgRNA. Standard deviations of the results from 4 replicates are shown. Red arrow indicates fold change compared to control.

Classic studies of Z. mobilis physiology (27) and contemporary work using strains with both hpn+ and hpn::Tn alleles (28) have linked hopanoid production and ethanol concentration or resistance; however, it is unknown whether hopanoids provide resistance to advanced biofuels, such as IBA. To examine the relationship between hpn and shc genes, we first determined the concentration of IBA needed to partially inhibit growth of Z. mobilis in sealed 96-well deep-well plates. We found that the addition of 1.25% (vol/vol) IBA to rich medium inhibited Z. mobilis growth by ∼50% (Fig. S6). We then grew our CRISPRi strains targeting hpn-shc operon genes with a subsaturating concentration of inducer (100 μM IPTG) for a limited number of generations in the presence or absence of IBA. Under these conditions, the strain in which hpnH was targeted was the only knockdown tested that showed increased sensitivity to IBA, with a 2.9-fold reduction in the final optical density at 600 nm (OD600) at the end of the growth period relative to that of a nontargeting control sgRNA strain (Fig. 4C). HpnH performs the first enzymatic step in hopanoid side chain synthesis (Fig. S5) (59), suggesting that buildup of the core hopanoids diploptene and diplopterol may compromise IBA tolerance in Z. mobilis. We conclude that hopanoid biosynthesis operons are essential in Z. mobilis but knockdown of extended hopanoids does not alter IBA tolerance.

DISCUSSION

The lack of genetic tools for bioenergy-relevant, nonmodel bacteria has slowed progress toward engineering efficient production strains for advanced biofuels, such as IBA. Our optimized CRISPRi system for Z. mobilis overcomes this obstacle by enabling programmable, inducible, and titratable control of overexpression of both nonessential and essential genes. Because Z. mobilis CRISPRi also functions well in the distantly related bacterium E. coli, the system as constructed may have broad utility across species. Our path to optimizing Mobile-CRISPRi for Z. mobilis revealed valuable lessons that may be generalizable across species: first, identify limiting components (either dCas9 or sgRNAs) by overexpression, and second, design strong synthetic promoters that take advantage of conserved interactions between RNA polymerase holoenzyme and DNA (48, 49) to improve portability.

CRISPRi is the ideal genetic tool to explore the unusual genetics of Z. mobilis. Because many metabolic genes are predicted to be essential (60), controlling metabolic flux may require constructing strains with partial knockdowns of essential genes. Mismatch-CRISPRi libraries are particularly well suited for empirically defining relationships between knockdown of metabolic genes (and associated changes in flux) and fitness, as strains comprising knockdown gradients of metabolic genes can be pooled and tested under a variety of growth conditions, with fitness measured by next-generation sequencing of sgRNA spacers. Furthermore, the Z. mobilis chromosome is possibly polyploid (28, 29), or at least capable of duplicating at high frequency; this can cause problems with deletion/transposon insertion analysis of essential genes or other genes that have a strong impact on fitness. For instance, a high-throughput analysis of isolated transposon insertion mutants revealed that there was no significant difference in the probability of a transposon inserting into a predicted essential versus nonessential gene (29), suggesting a polyploid chromosome and underscoring issues with interpreting insertion/deletion results in Z. mobilis. In contrast, CRISPRi is largely unaffected by polyploidy—it is capable of targeting essential genes across multiple copies of the chromosome as long as the sgRNA-dCas9 complex is expressed at high enough levels to account for additional targets.

Numerous studies have linked hopanoid production and ethanol tolerance in Z. mobilis (27, 30, 61), but whether hopanoids provide resistance to nonphysiological alcohols, such as IBA, remains unclear. Our CRISPRi results suggest that wild-type levels of extended hopanoids do not impart IBA resistance, and instead, that preventing synthesis of extended (C35) hopanoids by blocking hpnH expression causes sensitivity. The simplest explanation for these results is that the core C30 hopanoids, diploptene and diplopterol, accumulate in the cell and negatively impact the outer membrane. Further evaluation of this hypothesis will require careful tracking of the relationships between knockdown extent, fitness, IBA concentration, and levels of individual hopanoid species.

CRISPRi enables essential gene phenotyping, but we currently lack a unified standard for designating genes as “essential” or “nonessential” using CRISPRi. The ability of any CRISPRi system to identify essential genes depends on repression efficiency, the number of generations cells are grown after knockdown is induced, and the frequency at which CRISPRi suppressors exist within the cell population. High-efficiency systems (arbitrarily defined as >50-fold knockdown) are more likely to approximate null mutations for all queried genes, while low-efficiency or partially induced systems (<10-fold knockdown) will only report essentiality for a subset of genes that are most sensitive to knockdown (40); therefore, we recommend using high-efficiency systems for genome-wide analysis of essentiality using CRISPRi. For experiments using model bacteria with rigorously defined essential gene sets (e.g., E. coli [62] and B. subtilis [63]), the extent of CRISPRi knockdown and number of generations grown postknockdown can be calibrated to maximize recovery of essential genes (64). For nonmodel bacteria, such as Z. mobilis, we recommend comparing the viability or fitness of query genes to that of universally conserved essential genes (e.g., core ribosomal proteins). We further recommend removing repetitive sequences from CRISPRi constructs to reduce the likelihood that homologous recombination events inactivate CRISPRi and contribute to the assay background.

Our optimized CRISPRi system opens the door to high-throughput, systematic analysis of gene function in Z. mobilis. We envision that such screens will be invaluable for identifying genes involved in resistance to hydrolysate or biofuel inhibitors, genetic fingerprinting of hydrolysates from different plant sources or environments, and improving our understanding of the unique metabolism of Z. mobilis. We anticipate that this information will power the next generation of biofuel production strains, resulting in higher yields of advanced biofuels and bioproducts.

MATERIALS AND METHODS

Strains and growth conditions.

Strains are listed in Table 1. Escherichia coli was grown in Lennox LB broth (10 g tryptone, 5 g yeast extract, 5 g NaCl per liter; BD 240230) at 37°C aerobically in a flask with shaking at 250 rpm, in a culture tube on a roller drum, or in a deep 96-well plate with shaking at 900 rpm. Zymomonas mobilis was grown in DSMZ medium 10 (DSMZ10; 10 g peptone and 10 g yeast extract per liter plus 2% glucose) at 30°C aerobically without shaking. The medium was solidified with 1.5% agar for growth on plates. Antibiotics were added when necessary—for E. coli, 100 μg/ml ampicillin, 20 μg/ml chloramphenicol, or 30 μg/ml kanamycin, and for Z. mobilis, 100 μg/ml chloramphenicol, or 120 μg/ml kanamycin. Diaminopimelic acid (DAP) was added at 300 μM to support growth of dap-negative E. coli strains. IPTG (isopropyl β-d-1-thiogalactopyranoside) at 0.1 to 1 mM was added where indicated in the figures or figure legends. All strains were preserved in 15% glycerol at −80°C.

Plasmid construction.

Plasmids and construction details are listed in Table 2, and a representative plasmid map is shown in Fig. S7 in the supplemental material; oligonucleotides and synthetic DNA are listed in Table 3. pir-dependent plasmids were propagated in E. coli strain BW25141 (sJMP146) and other plasmids in E. coli strain DH10B (sJMP032). Plasmids were assembled from fragments (linearized vector, PCR products, and/or synthetic DNA) using the NEBuilder hifi DNA assembly kit (catalog number E2621; New England Biolabs [NEB]). Plasmids were cut with restriction enzymes from NEB. Linearized plasmids were re-ligated using T4 DNA ligase (catalog number M0202; NEB). Fragments were amplified using Q5 DNA polymerase (catalog number M0491; NEB), followed by digestion with DpnI. Fragments were purified using the Monarch PCR & DNA cleanup kit (catalog number T1030; NEB) or the Zymo Research DNA Clean & Concentrator-5 kit (catalog number D4004) after digestion or amplification. Plasmids were transformed into electrocompetent E. coli cells using a Bio-Rad Gene Pulser Xcell on the EC1 setting. Plasmids were purified using the GeneJet plasmid miniprep kit (catalog number K0503; Thermo Scientific) or the PureLink HiPure Plasmid Midiprep kit (catalog number K210005; Invitrogen). Site-directed mutagenesis of plasmids was performed by DNA synthesis with 2.5 U PfuUltra II fusion HS DNA polymerase (Agilent), 0.2 μM single oligonucleotide bearing the change, 0.2 mM deoxynucleoside triphosphates (dNTPs), and 50 ng plasmid DNA in a 25-μl reaction mixture with a 1-min/kb extension time at 68°C, followed by DpnI digestion. sgRNA-encoding sequences were cloned into CRISPRi plasmids between the BsaI sites with inserts prepared by one of two methods. In method one, two 24-nucleotide (nt) oligonucleotides were designed to overlap such that when annealed, their ends would be complementary to the BsaI-cut ends on the vector. Oligonucleotides (2 μM each) were annealed in 1× CutSmart buffer (NEB) at 95°C for 5 min, followed by cooling to room temperature. For method two, fragments were amplified by PCR with primers oJMP197 and oJMP198 from a 78-nt oligonucleotide, followed by digestion with BsaI-HF-v2 (catalog number R3733; NEB) and purification with the Monarch DNA purification kit (NEB) following the manufacturer’s oligonucleotide purification protocol. Inserts (2 μl of a 1:40 dilution of annealed oligonucleotides or 2 ng purified digested PCR product) were ligated into 50 ng BsaI-digested vector. Oligonucleotides and synthetic DNA gBlocks were purchased from Integrated DNA Technologies (Coralville, IA). Sequencing was performed by Functional Biosciences (Madison, WI).

TABLE 2.

Plasmids

Plasmid Descriptiona Construction/notes Marker(s)b Reference or source
pJMP445 pRL814 Broad-host-range plasmid, pBBR1 ori, sfGFP Specr 67
pJMP1039 pTn7C1 Tn7 transposase expression Ampr 41
pJMP1183 pTn7C89.1 Mobile-CRISPRi RFP test plasmid (RR1 sgRNA) Ampr, Kanr 41
pJMP1185 pTn7C90.1 Mobile-CRISPRi RFP test plasmid (nontargeting sgRNA) Ampr, Kanr 41
pJMP1337 MCi (ICE::CRISPRi, Spy dCas9) Ampr, Kanr 41
pJMP1339 MCi (Tn7::CRISPRi, Hsa dCas9) Ampr, Kanr 41
pJMP1356 MCi (Hsa dCas9) Ampr, Cmr 41
pJMP2030 pRL814 (sfGFP ΔXhoI) Eliminate XhoI site from sfGFP gene in pRL814 (pJMP445) by site-directed mutagenesis with oJMP076 Specr This study
pJMP2044 MCi (Cmr) (Spy dCas9) pJMP1356 cut with AscI and SpeI, assembled with Spy dCas9, amplified from pJMP1337 with oJMP072 and oJMP073 Ampr, Cmr This study
pJMP2046 MCi-vA-GFP_BsaI pJMP2044 cut with EcoRI and PmeI, assembled with gBlock oJMP079 and sfGFP (no XhoI), amplified from pJMP2030 with oJMP074 and oJMP075 Ampr, Cmr This study
pJMP2048 MCi-vA-GFP_gfp pJMP2044 cut with EcoRI and PmeI, assembled with gBlock oJMP080 and sfGFP (no XhoI), amplified from pJMP2030 with oJMP074 and oJMP075 Ampr, Cmr This study
pJMP2093 MCi-vB-GFP_BsaI pJMP2046 cut with EcoRI, assembled with gBlock oJMP191 Ampr, Cmr This study
pJMP2095 MCi-vB-GFP_gfp pJMP2048 cut with EcoRI, assembled with gBlock oJMP192 Ampr, Cmr This study
pJMP2132 MCi-vB_BsaI pJMP2093 cut with PmeI and re-ligated to remove sfGFP Ampr, Cmr This study
pJMP2316 pSRK-kan Broad-host-range plasmid, pBBR1 ori Kanr 68
pJMP2317 pSRK-kan-sgRNA Vector backbone amplified from pJMP2316 with oJMP313 and oJMP314, assembled with sgRNA cassette, amplified from pJMP2095 with oJMP315 and oJMP316 Kanr This study
pJMP2319 pSRK-kan-dCas9 Vector backbone amplified from pJMP2316 with oJMP313 and oJMP314, assembled with dCas9 cassette amplified from pJMP2095 with oJMP317 and oJMP318 Kanr This study
pJMP2367 MCi-vC-GFP_BsaI Assemble EcoRI-cut pJMP2093 with gBlock oJMP347 Ampr, Cmr This study
pJMP2369 MCi-vD-GFP_BsaI Assemble EcoRI-cut pJMP2093 with gBlock oJMP348 Ampr, Cmr This study
pJMP2371 MCi-vE-GFP_BsaI Assemble EcoRI-cut pJMP2093 with gBlock oJMP349 Ampr, Cmr This study
pJMP2373 MCi-vF-GFP_BsaI Assemble EcoRI-cut pJMP2093 with gBlock oJMP350 Ampr, Cmr This study
pJMP2375 MCi-vC-GFP_gfp Assemble EcoRI-cut pJMP2093 with gBlock oJMP351 Ampr, Cmr This study
pJMP2377 MCi-vD-GFP_gfp Assemble EcoRI-cut pJMP2093 with gBlock oJMP352 Ampr, Cmr This study
pJMP2379 MCi-vE-GFP_gfp Assemble EcoRI-cut pJMP2093 with gBlock oJMP353 Ampr, Cmr This study
pJMP2381 MCi-vF-GFP_gfp Assemble EcoRI-cut pJMP2093 with gBlock oJMP354 Ampr, Cmr This study
pJMP2391 MCi-vB_hpnC-1 (ZMO869) Annealed oJMP355 and oJMP356 ligated into BsaI-cut pJMP2132 Ampr, Cmr This study
pJMP2393 MCi-vB_hpnC-2 (ZMO869) Annealed oJMP357 and oJMP358 ligated into BsaI-cut pJMP2132 Ampr, Cmr This study
pJMP2395 MCi-vB_hpnF-1 (ZMO872) Annealed oJMP359 and oJMP360 ligated into BsaI-cut pJMP2132 Ampr, Cmr This study
pJMP2397 MCi-vB_hpnF-2 (ZMO872) Annealed oJMP361 and oJMP362 ligated into BsaI-cut pJMP2132 Ampr, Cmr This study
pJMP2399 MCi-vB_hpnH-1 (ZMO874) Annealed oJMP363 and oJMP364 ligated into BsaI-cut pJMP2132 Ampr, Cmr This study
pJMP2401 MCi-vB_hpnH-2 (ZMO874) Annealed oJMP365 and oJMP366 ligated into BsaI-cut pJMP2132 Ampr, Cmr This study
pJMP2403 MCi-vB_hpnI-1 (ZMO972) Annealed oJMP367 and oJMP368 ligated into BsaI-cut pJMP2132 Ampr, Cmr This study
pJMP2405 MCi-vB_hpnI-2 (ZMO972) Annealed oJMP369 and oJMP370 ligated into BsaI-cut pJMP2132 Ampr, Cmr This study
pJMP2407 MCi-vB_shc2-1 (ZMO1548) Annealed oJMP371 and oJMP372 ligated into BsaI-cut pJMP2132 Ampr, Cmr This study
pJMP2409 MCi-vB_shc2-1 (ZMO1548) Annealed oJMP373 and oJMP374 ligated into BsaI-cut pJMP2132 Ampr, Cmr This study
pJMP2415 MCi-vB_rfp Annealed oJMP003 and oJMP021 ligated into BsaI-cut pJMP2132 Ampr, Cmr This study
pJMP2480 MCi-vC_BsaI pJMP2367 cut with PmeI and re-ligated to remove sfGFP Ampr, Cmr This study
pJMP2509 MCi-vC-GFP_gmc1 Annealed oJMP400 and oJMP401 ligated into BsaI-cut pJMP2480 Ampr, Cmr This study
pJMP2511 MCi-vC-GFP_gmc2 Annealed oJMP402 and oJMP403 ligated into BsaI-cut pJMP2480 Ampr, Cmr This study
pJMP2513 MCi-vC-GFP_gmc3 Annealed oJMP404 and oJMP405 ligated into BsaI-cut pJMP2480 Ampr, Cmr This study
pJMP2515 MCi-vC-GFP_gmc4 Annealed oJMP406 and oJMP407 ligated into BsaI-cut pJMP2480 Ampr, Cmr This study
pJMP2517 MCi-vC-GFP_gmc5 Annealed oJMP408 and oJMP409 ligated into BsaI-cut pJMP2480 Ampr, Cmr This study
pJMP2519 MCi-vC-GFP_gmc6 Annealed oJMP410 and oJMP411 ligated into BsaI-cut pJMP2480 Ampr, Cmr This study
pJMP2521 MCi-vC-GFP_gmc7 Annealed oJMP412 and oJMP413 ligated into BsaI-cut pJMP2480 Ampr, Cmr This study
pJMP2523 MCi-vC-GFP_gmc8 Annealed oJMP414 and oJMP415 ligated into BsaI-cut pJMP2480 Ampr, Cmr This study
pJMP2525 MCi-vC-GFP_gmc9 Annealed oJMP416 and oJMP417 ligated into BsaI-cut pJMP2480 Ampr, Cmr This study
pJMP2597 MCi-vC-pdc-1 (ZMO1360) Annealed oJMP467 and oJMP468 ligated into BsaI-cut pJMP2480 Ampr, Cmr This study
pJMP2598 MCi-vC-pdc-3 (ZMO1360) Annealed oJMP469 and oJMP470 ligated into BsaI-cut pJMP2480 Ampr, Cmr This study
pJMP2599 MCi-vC-rplL (ZMO0728) Amplify from opZ1-1-13 with oJMP197 and oJMP198, digest fragment with BsaI, and ligate into BsaI-cut pJMP2480 Ampr, Cmr This study
a

MCi, Mobile-CRISPRi; ICE, integrative conjugative element; Hsa, Homo sapiens codon-optimized dCas9; vA to vF, sgRNA promoter variants; RFP, red fluorescent protein; GFP, green fluorescent protein; BsaI, vector with BsaI cloning site for sgRNA.

b

Cmr, chloramphenicol resistance; Ampr, ampicillin resistance, Kanr, kanamycin resistance.

TABLE 3.

Oligonucleotides and synthetic DNA

Oligonucleotide Sequence (5′→3′) Description Usage
oJMP003 TAGTAACTTTCAGTTTAGCGGTCT rfp_T pJMP2415
oJMP021 AAACAGACCGCTAAACTGAAAGTT rfp_B pJMP2415
oJMP057 CCAAGGTGCATCCTCTCATT Zmo_Tn7_check_A
oJMP058 TATCGGACAATCGGGAAGAC Zmo_Tn7_check_B
oJMP059 GCCCCGATCGTCTATGCTAT Zmo_Tn7_check_C
oJMP060 CGCCCCTCTTTAATACGACG Tn7R_check
oJMP072 CGCTTTTTTTACGTCTGCAGACTAGTAAAATTTATCAAAAAGAGTGTTGACTTGTGAGCGGATAACAATGATACTTAGATTCAATTGTGAGCGGATAACAATTGAGCGAGAAGGAGGACTAGTATGGATAAGAAATACTCAATAGGC T7A1_O3O4-dcas9_F pJMP2044
oJMP073 TTTGGTACCGAGGCTGCAA T7A1_O3O4-dcas9_R pJMP2044
oJMP074 GGAGAAGAACTTTTCACTGGAGT sfgfp_F pJMP2046, pJMP2048
oJMP075 GCAAATCCAGGAGGTCGTTTAAACTTATTATTTGTAGAGCTCATCCATGCCATGTG sfgfp_R pJMP2046, pJMP2048
oJMP076 ATGGAAACATTCTTGGACACAAACTGGAGTACAACTTTAACTCACACAATG sfgfp_no_XhoI_QC pJMP2030
oJMP079 ACCTATCGACTGAGCTGAAAGAATTCGCTCACTCATTAGGCACCCCAGGCTTTACACTTTATGCTTCCGGCTCGTATAATGTCTAGTTGAGACCAACTTTGGTCTCCACCATAGCGGTCGGTCTCTGTTTAAGAGCTATGCTGGAAACAGCATAGCAAGTTTAAATAAGGCTAGTCCGTTATCAACTTGAAAAAGTGGCACCGAGTCGGTGCTTTTTTTTTGAATTCATGTGGCTGACCGTTCTGTTGTCTCTCGCTCTTCCGAGTAGACGAACAATAAGGCCTCCCTAACGGGGGGCCTTTTTTATTGATAACAAAAGTCAGTGCTTCCGCTATTTCCAAAATACCGGGCTAATACGGTTTAAACGAAAATTTATCAAAAAGAGTATTGACTTAAAGTCTAACCTATAGGATACTTACAGCCAGATCTGAGCGAGAAGGAGGTAAAGTATGAGCAAAGGAGAAGAACTTTTCACTGG Zmo_M-Ci_gBlock_BsaI pJMP2046
oJMP080 ACCTATCGACTGAGCTGAAAGAATTCGCTCACTCATTAGGCACCCCAGGCTTTACACTTTATGCTTCCGGCTCGTATAATGTCTAGTCATCTAATTCAACAAGAATTGTTTAAGAGCTATGCTGGAAACAGCATAGCAAGTTTAAATAAGGCTAGTCCGTTATCAACTTGAAAAAGTGGCACCGAGTCGGTGCTTTTTTTTTGAATTCATGTGGCTGACCGTTCTGTTGTCTCTCGCTCTTCCGAGTAGACGAACAATAAGGCCTCCCTAACGGGGGGCCTTTTTTATTGATAACAAAAGTCAGTGCTTCCGCTATTTCCAAAATACCGGGCTAATACGGTTTAAACGAAAATTTATCAAAAAGAGTATTGACTTAAAGTCTAACCTATAGGATACTTACAGCCAGATCTGAGCGAGAAGGAGGTAAAGTATGAGCAAAGGAGAAGAACTTTTCACTGG Zmo_M-Ci_gBlock_GR1 pJMP2048
oJMP191 ACCTATCGACTGAGCTGAAAGAATTCGCTCACTCATTAGGCACCCCAGGCTTTACAATTGTGAGCGCTCACAATTATAATGTCTAGTTGAGACCAACTTTGGTCTCCACCATAGCGGTCGGTCTCTGTTTAAGAGCTATGCTGGAAACAGCATAGCAAGTTTAAATAAGGCTAGTCCGTTATCAACTTGAAAAAGTGGCACCGAGTCGGTGCTTTTTTTTTGAATTCATGTGGCTGACCGTTCTGTT BsaI_gBlock pJMP2093
oJMP192 ACCTATCGACTGAGCTGAAAGAATTCGCTCACTCATTAGGCACCCCAGGCTTTACAATTGTGAGCGCTCACAATTATAATGTCTAGTCATCTAATTCAACAAGAATTGTTTAAGAGCTATGCTGGAAACAGCATAGCAAGTTTAAATAAGGCTAGTCCGTTATCAACTTGAAAAAGTGGCACCGAGTCGGTGCTTTTTTTTTGAATTCATGTGGCTGACCGTTCTGTT GR1_gblock pJMP2095
oJMP197 GTCTATTGAAGTACCTGC Z1_amplify_A pJMP2599
oJMP198 CGATGCTCCTCCAAGATA Z1_amplify_B pJMP2599
oJMP313 GCTAGCAATTCGAAAGCAAATTCGACCC pSRK-V-F pJMP2317, pJMP2319
oJMP314 ATTGCGTTGCGCTCACTGCC pSRK-V-R pJMP2317, pJMP2319
oJMP315 GGCAGTGAGCGCAACGCAATACCTATCGACTGAGCTGAAAGAAT sgRNA-F pJMP2317
oJMP316 GGTCGAATTTGCTTTCGAATTGCTAGCAACAGAACGGTCAGCCACAT sgRNA-R pJMP2317
oJMP317 GGCAGTGAGCGCAACGCAATTGCAGACTAGTAAAATTTATCAAAAAGAGTGTTGAC dCas9-F pJMP2319
oJMP318 GGTCGAATTTGCTTTCGAATTGCTAGCGGCGCGCCTTATTACAGATCTTC dCas9-R pJMP2319
oJMP347 ACCTATCGACTGAGCTGAAAGAATTCGGAAAATTTTTTTTCAAAAGTACTTGAAATTGTGAGCGCTCACAATTATAATTCTAGTAGAGACCAACTTTGGTCTCCACCATAGCGGTCGGTCTCTGTTTAAGAGCTATGCTGGAAACAGCATAGCAAGTTTAAATAAGGCTAGTCCGTTATCAACTTGAAAAAGTGGCACCGAGTCGGTGCTTTTTTTTTGAATTCATGTGGCTGACCGTTCTGTT C_gBlock pJMP2367
oJMP348 ACCTATCGACTGAGCTGAAAGAATTCGGAAAATTTTTTTTCAAAAGTACTTGAATTGTGAGCGGATAACAATTATAATTCTAGTAGAGACCAACTTTGGTCTCCACCATAGCGGTCGGTCTCTGTTTAAGAGCTATGCTGGAAACAGCATAGCAAGTTTAAATAAGGCTAGTCCGTTATCAACTTGAAAAAGTGGCACCGAGTCGGTGCTTTTTTTTTGAATTCATGTGGCTGACCGTTCTGTT D_gBlock pJMP2369
oJMP349 ACCTATCGACTGAGCTGAAAGAATTCGGAAAATTTTTTTTCAAAAGTACTTTACAATTGTGAGCGCTCACAATTATAATTCTAGTAGAGACCAACTTTGGTCTCCACCATAGCGGTCGGTCTCTGTTTAAGAGCTATGCTGGAAACAGCATAGCAAGTTTAAATAAGGCTAGTCCGTTATCAACTTGAAAAAGTGGCACCGAGTCGGTGCTTTTTTTTTGAATTCATGTGGCTGACCGTTCTGTT E_gBlock pJMP2371
oJMP350 ACCTATCGACTGAGCTGAAAGAATTCGGAAAATTTTTTTTCAAAAGTACTTTAAATTGTGAGCGGATAACAATTATAATTCTAGTAGAGACCAACTTTGGTCTCCACCATAGCGGTCGGTCTCTGTTTAAGAGCTATGCTGGAAACAGCATAGCAAGTTTAAATAAGGCTAGTCCGTTATCAACTTGAAAAAGTGGCACCGAGTCGGTGCTTTTTTTTTGAATTCATGTGGCTGACCGTTCTGTT F_gBlock pJMP2373
oJMP351 ACCTATCGACTGAGCTGAAAGAATTCGGAAAATTTTTTTTCAAAAGTACTTGAAATTGTGAGCGCTCACAATTATAATTCTAGTCATCTAATTCAACAAGAATTGTTTAAGAGCTATGCTGGAAACAGCATAGCAAGTTTAAATAAGGCTAGTCCGTTATCAACTTGAAAAAGTGGCACCGAGTCGGTGCTTTTTTTTTGAATTCATGTGGCTGACCGTTCTGTT C-gfp_gBlock pJMP2375
oJMP352 ACCTATCGACTGAGCTGAAAGAATTCGGAAAATTTTTTTTCAAAAGTACTTGAATTGTGAGCGGATAACAATTATAATTCTAGTCATCTAATTCAACAAGAATTGTTTAAGAGCTATGCTGGAAACAGCATAGCAAGTTTAAATAAGGCTAGTCCGTTATCAACTTGAAAAAGTGGCACCGAGTCGGTGCTTTTTTTTTGAATTCATGTGGCTGACCGTTCTGTT D-gfp_gBlock pJMP2377
oJMP353 ACCTATCGACTGAGCTGAAAGAATTCGGAAAATTTTTTTTCAAAAGTACTTTACAATTGTGAGCGCTCACAATTATAATTCTAGTCATCTAATTCAACAAGAATTGTTTAAGAGCTATGCTGGAAACAGCATAGCAAGTTTAAATAAGGCTAGTCCGTTATCAACTTGAAAAAGTGGCACCGAGTCGGTGCTTTTTTTTTGAATTCATGTGGCTGACCGTTCTGTT E-gfp_gBlock pJMP2379
oJMP354 ACCTATCGACTGAGCTGAAAGAATTCGGAAAATTTTTTTTCAAAAGTACTTTAAATTGTGAGCGGATAACAATTATAATTCTAGTCATCTAATTCAACAAGAATTGTTTAAGAGCTATGCTGGAAACAGCATAGCAAGTTTAAATAAGGCTAGTCCGTTATCAACTTGAAAAAGTGGCACCGAGTCGGTGCTTTTTTTTTGAATTCATGTGGCTGACCGTTCTGTT F-gfp_gBlock pJMP2381
oJMP355 TAGTTCCTTTTCCAGAAACCAAAG zmo_hpnC_A_T pJMP2391
oJMP356 AAACCTTTGGTTTCTGGAAAAGGA zmo_hpnC_A_B pJMP2391
oJMP357 TAGTATAATAATAGGCCGATATTC zmo_hpnC_B_T pJMP2393
oJMP358 AAACGAATATCGGCCTATTATTAT zmo_hpnC_B_B pJMP2393
oJMP359 TAGTCGGGCTATGATGAAAAGCCG zmo_hpnF_A_T pJMP2395
oJMP360 AAACCGGCTTTTCATCATAGCCCG zmo_hpnF_A_B pJMP2395
oJMP361 TAGTCGGGTGGCCTTTTGGATAAT zmo_hpnF_B_T pJMP2397
oJMP362 AAACATTATCCAAAAGGCCACCCG zmo_hpnF_B_B pJMP2397
oJMP363 TAGTATCCGTAAAACCTGACTAAA zmo_hpnH_A_T pJMP2399
oJMP364 AAACTTTAGTCAGGTTTTACGGAT zmo_hpnH_A_B pJMP2399
oJMP365 TAGTGGTTCAAGCATCAAAACCAG zmo_hpnH_B_T pJMP2401
oJMP366 AAACCTGGTTTTGATGCTTGAACC zmo_hpnH_B_B pJMP2401
oJMP367 TAGTTGTCAAAAGGACATGCAGGA zmo_hpnI_A_T pJMP2403
oJMP368 AAACTCCTGCATGTCCTTTTGACA zmo_hpnI_A_B pJMP2403
oJMP369 TAGTCCACCACAGCACGACAATCG zmo_hpnI_B_T pJMP2405
oJMP370 AAACCGATTGTCGTGCTGTGGTGG zmo_hpnI_B_B pJMP2405
oJMP371 TAGTATCGGAAGCCGGTTTATACG zmo_shc2_A_T pJMP2407
oJMP372 AAACCGTATAAACCGGCTTCCGAT zmo_shc2_A_B pJMP2407
oJMP373 TAGTTTGCCGTCGGTTCAGGCCGC zmo_shc2_B_T pJMP2409
oJMP374 AAACGCGGCCTGAACCGACGGCAA zmo_shc2_B_B pJMP2409
oJMP400 TAGTTTCCGTTGGGATCTTTCGAA gmc1_T pJMP2409
oJMP401 AAACTTCGAAAGATCCCAACGGAA gmc1_B pJMP2409
oJMP402 TAGTTAGTACATAACCTTCGGGCA gmc2_T pJMP2411
oJMP403 AAACTGCCCGAAGGTTATGTACTA gmc2_B pJMP2411
oJMP404 TAGTGTCAGAGTAGTGTCAAGTGT gmc3_T pJMP2413
oJMP405 AAACACACTTGACACTACTCTGAC gmc3_B pJMP2413
oJMP406 TAGTGGCAAAGCATTGAAAACCAT gmc4_T pJMP2415
oJMP407 AAACATGGTTTTCAATGCTTTGCC gmc4_B pJMP2415
oJMP408 TAGTGCGTTCCTGTACATAACCCT gmc5_T pJMP2417
oJMP409 AAACAGGGTTATGTACAGGAACGC gmc5_B pJMP2417
oJMP410 TAGTCATCTAATTCAACAAGAATT gmc6_T pJMP2419
oJMP411 AAACAATTCTTGTTGAATTAGATG gmc6_B pJMP2419
oJMP412 TAGTATGTTGTCACGCTTTTCGTT gmc7_T pJMP2421
oJMP413 AAACAACGAAAAGCGTGACAACAT gmc7_B pJMP2421
oJMP414 TAGTAGTAGTGCAAAGAAATTTAA gmc8_T pJMP2423
oJMP415 AAACTTAAATTTCTTTGCACTACT gmc8_B pJMP2423
oJMP416 TAGTAAACGACAGATTGTGTCGAC gmc9_T pJMP2425
oJMP417 AAACGTCGACACAATCTGTCGTTT gmc9_B pJMP2425
oJMP467 TAGTGACAAGCCGCTCCGCTAAAT zmo1360-pdc-1-T pJMP2597
oJMP468 AAACATTTAGCGGAGCGGCTTGTC zmo1360-pdc-1-B pJMP2597
oJMP469 TAGTGACGGCTGCTGCTGCGCCTT zmo1360-pdc-3-T pJMP2598
oJMP470 AAACAAGGCGCAGCAGCAGCCGTC zmo1360-pdc-3-B pJMP2598
opZ1-1-13 GTCTATTGAAGTACCTGCGGTCTCTTAGTAAGTTCAGCTGCTTCAAGAAGTTTAGAGACCTATCTTGGAGGAGCATCG opZ1-1-13-zmo0728_rplL pJMP2599

Transfer of CRISPRi system to E. coli and Z. mobilis.

Strains with a chromosomally located CRISPRi expression cassette were constructed by triparental mating of two donor strains—one with a plasmid encoding Tn7 transposase and another with a plasmid containing a Tn7 transposon encoding the CRISPRi system—and a recipient strain (either E. coli BW25113 or Z. mobilis ZM4 [PK15436]). The method was as described in Peters et al. (41), with several modifications. Briefly, all matings used E. coli WM6026, which is pir+ to support pir-dependent plasmid replication, dap negative, making it dependent on diaminopimelic acid (DAP) for growth, and encodes the RP4 transfer machinery required for conjugation. Donor strains were grown for ∼16 h at 37°C in LB plus 100 μg/ml ampicillin and 300 μM DAP. The E. coli recipient was grown for ∼16 h at 37°C in LB. The Z. mobilis recipient was grown for ∼24 to 30 h at 30°C in DSMZ10. Cells were centrifuged at 4,000 × g for 5 min and gently resuspended twice in an equal volume of fresh medium with no antibiotic or DAP. For E. coli recipients, 700 μl LB and 100 μl each donor and recipient were mixed in a sterile 1.5-ml microcentrifuge tube. For Z. mobilis recipients, 200 μl DSMZ10, 500 μl Z. mobilis, 200 μl transposase donor, and 100 μl transposon donor were mixed in a sterile 1.5-ml microcentrifuge tube. Cells were centrifuged at 4,000 × g for 3 min, gently resuspended in 25 μl LB or DSMZ10, and pipetted onto a 13-mm cellulose filter (number HAWG01300; MF-Millipore) placed on a prewarmed agar plate (LB for E. coli or DSMZ10 for Z. mobilis). Plates were incubated at 37°C for 2 to 6 h for E. coli and 30°C for 24 h for Z. mobilis. After the incubation period, using sterile forceps, filters were placed into sterile 1.5-ml microcentrifuge tubes containing 200 μl sterile 1× phosphate-buffered saline (PBS), vortexed for 20 s to dislodge cells from filters, diluted in 1× PBS, and plated on appropriate medium for the recipient, with antibiotic to select for the transposon (see above) and no DAP (to select against the donor). The efficiency of transposition was generally ∼1 in 103 for E. coli and ∼1 in 105 to 106 for Z. mobilis. Isolated colonies were generally obtained from ∼10 to 100 μl of a 1:100 dilution per plate, and isolated colonies were restruck for isolation to ensure purity.

CRISPRi insertion onto Z. mobilis chromosome.

Insertion of the CRISPRi expression cassette into the Tn7 att site downstream from glmS in Z. mobilis was confirmed by PCR with primers oJMP057 and oJMP058 (flanking the insertion site) and oJMP059 and oJMP060 (upstream from the insertion site and within the CRISPRi transposon).

CRISPRi stability in Z. mobilis.

Z. mobilis strains with chromosomally located CRISPRi expression cassettes (6 individual isolates) were grown in liquid culture medium with antibiotic selection to saturation. This culture was serially diluted to 10−5 into nonselective medium (starting OD600 of ∼0.00002) and grown for ∼17 generations back to saturation (OD600 of ∼2.0). Dilution and growth were repeated 2 additional times for a total of ∼50 generations prior to plating on nonselective plates. Forty-eight isolated colonies were selected and patched on selective and nonselective plates, and all strains retained the ability to grow on the antibiotic to which resistance was conferred by the chromosomally located CRISPRi expression cassette.

GFP and RFP knockdown assays.

GFP or RFP knockdown was measured using a plate reader (Tecan Infinite 200 Pro M Plex). Cell density was determined by OD600, and fluorescence was measured by excitation/emission at 482/515 nm for GFP and 584/607 nm for RFP. Initial cultures (n = 4) were grown from single colonies to saturation (∼30 h for Z. mobilis or ∼16 h for E. coli) in 1 ml medium in deep 96-well plates. These cultures were serially diluted 1:1,000 (Z. mobilis) or 1:10,000 (E. coli) into 1 ml fresh medium (no antibiotic and 0 to 1 mM IPTG as indicated) and grown back to saturation (∼24 to 30 h for Z. mobilis or ∼8 to 16 h for E. coli). Pelleted cells were resuspended in 1 ml 1× PBS and, if necessary, diluted, and 200 μl was transferred to a clear-bottom black microtiter plate and measured in the plate reader as indicated above. Fluorescence values were normalized to cell density and to measurements from strains not expressing GFP.

Gene knockdown spot dilution assay.

Z. mobilis strains with chromosomally located CRISPRi expression cassettes (2 individual isolates) were grown to saturation in liquid culture medium with antibiotic selection. These cultures were serially diluted 1:10 in nonselective medium, and 3-μl amounts were spotted onto plates containing 0, 0.1 mM, or 1 mM IPTG that were incubated at 30°C aerobically prior to analysis.

Gene knockdown growth assay.

Z. mobilis strains with chromosomally located CRISPRi expression cassettes (n = 2 individual isolates) were grown to saturation in liquid culture medium with antibiotic selection. These cultures were serially diluted 1:1,000 into nonselective medium with 0 or 0.1 mM IPTG and 0, 0.63%, 1.25%, or 2.5% isobutanol in a deep 96-well plate and incubated at 30°C aerobically prior to analysis of growth (measured as OD600).

Data availability.

Plasmids and their sequences are available from Addgene (Addgene identification numbers 160073 to 160080).

Supplementary Material

Supplemental file 1
AEM.01621-20-s0001.pdf (908.5KB, pdf)

ACKNOWLEDGMENTS

We thank members of the TerAvest, Landick, Kiley, Amador-Noguez, Reed, Sato, Hittinger and Gross labs for helpful feedback. We thank Yang Liu, Robert Landick, Piyush Lal, and Tricia Kiley for strains and plasmids. This work was supported by the U.S. Department of Energy, Office of Science, Office of Biological and Environmental Research, Great Lakes Bioenergy Research Center under Award Number DE-SC0018409.

Footnotes

Supplemental material is available online only.

REFERENCES

  • 1.He MX, Wu B, Qin H, Ruan ZY, Tan FR, Wang JL, Shui ZX, Dai LC, Zhu QL, Pan K, Tang XY, Wang WG, Hu QC. 2014. Zymomonas mobilis: a novel platform for future biorefineries. Biotechnol Biofuels 7:101. doi: 10.1186/1754-6834-7-101. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 2.Zhang K, Lu X, Li Y, Jiang X, Liu L, Wang H. 2019. New technologies provide more metabolic engineering strategies for bioethanol production in Zymomonas mobilis. Appl Microbiol Biotechnol 103:2087–2099. doi: 10.1007/s00253-019-09620-6. [DOI] [PubMed] [Google Scholar]
  • 3.Qiu M, Shen W, Yan X, He Q, Cai D, Chen S, Wei H, Knoshaug EP, Zhang M, Himmel ME, Yang S. 2020. Metabolic engineering of Zymomonas mobilis for anaerobic isobutanol production. Biotechnol Biofuels 13:15. doi: 10.1186/s13068-020-1654-x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 4.Wu B, Qin H, Yang Y, Duan G, Yang S, Xin F, Zhao C, Shao H, Wang Y, Zhu Q, Tan F, Hu G, He M. 2019. Engineered Zymomonas mobilis tolerant to acetic acid and low pH via multiplex atmospheric and room temperature plasma mutagenesis. Biotechnol Biofuels 12:10. doi: 10.1186/s13068-018-1348-9. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 5.Yang S, Fei Q, Zhang Y, Contreras LM, Utturkar SM, Brown SD, Himmel ME, Zhang M. 2016. Zymomonas mobilis as a model system for production of biofuels and biochemicals. Microb Biotechnol 9:699–717. doi: 10.1111/1751-7915.12408. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 6.Yang S, Vera JM, Grass J, Savvakis G, Moskvin OV, Yang Y, McIlwain SJ, Lyu Y, Zinonos I, Hebert AS, Coon JJ, Bates DM, Sato TK, Brown SD, Himmel ME, Zhang M, Landick R, Pappas KM, Zhang Y. 2018. Complete genome sequence and the expression pattern of plasmids of the model ethanologen Zymomonas mobilis ZM4 and its xylose-utilizing derivatives 8b and 2032. Biotechnol Biofuels 11:125. doi: 10.1186/s13068-018-1116-x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 7.Rogers PL, Lee KJ, Tribe DE. 1979. Kinetics of alcohol production by Zymomonas mobilis at high sugar concentrations. Biotechnol Lett 1:165–170. doi: 10.1007/BF01388142. [DOI] [Google Scholar]
  • 8.Rogers PL, Lee KJ, Skotnicki ML, Tribe DE. 1982. Ethanol production by Zymomonas mobilis. Adv Biochem Eng Biotechnol 23:37–84. [Google Scholar]
  • 9.Karsch T, Stahl U, Esser K. 1983. Ethanol production by Zymomonas and Saccharomyces, advantages and disadvantages. Eur J Appl Microbiol Biotechnol 18:387–391. doi: 10.1007/BF00504750. [DOI] [Google Scholar]
  • 10.Belauich JP, Senez JC. 1965. Influence of aeration and of pantothenate on growth yields of Zymomonas mobilis. J Bacteriol 89:1195–1200. doi: 10.1128/JB.89.5.1195-1200.1965. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 11.Choi YJ, Lee J, Jang Y-S, Lee SY. 2014. Metabolic engineering of microorganisms for the production of higher alcohols. mBio 5:e01524-14. doi: 10.1128/mBio.01524-14. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 12.Lee KY, Park JM, Kim TY, Yun H, Lee SY. 2010. The genome-scale metabolic network analysis of Zymomonas mobilis ZM4 explains physiological features and suggests ethanol and succinic acid production strategies. Microb Cell Fact 9:94. doi: 10.1186/1475-2859-9-94. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 13.Nouri H, Fouladiha H, Moghimi H, Marashi S-A. 2020. A reconciliation of genome-scale metabolic network model of Zymomonas mobilis ZM4. Sci Rep 10:7782. doi: 10.1038/s41598-020-64721-x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 14.Kalnenieks U, Pentjuss A, Rutkis R, Stalidzans E, Fell DA. 2014. Modeling of Zymomonas mobilis central metabolism for novel metabolic engineering strategies. Front Microbiol 5:42. doi: 10.3389/fmicb.2014.00042. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 15.Felczak MM, Jacobson TB, Ong WK, Amador-Noguez D, TerAvest MA. 2019. Expression of phosphofructokinase is not sufficient to enable Embden-Meyerhof-Parnas glycolysis in Zymomonas mobilis ZM4. Front Microbiol 10:2270. doi: 10.3389/fmicb.2019.02270. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 16.Conway T. 1992. The Entner-Doudoroff pathway: history, physiology and molecular biology. FEMS Microbiol Rev 9:1–27. doi: 10.1111/j.1574-6968.1992.tb05822.x. [DOI] [PubMed] [Google Scholar]
  • 17.Dawes EA, Ribbons DW, Large PJ. 1966. The route of ethanol formation in Zymomonas mobilis. Biochem J 98:795–803. doi: 10.1042/bj0980795. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 18.Jacobson TB, Adamczyk PA, Stevenson DM, Regner M, Ralph J, Reed JL, Amador-Noguez D. 2019. 2H and 13C metabolic flux analysis elucidates in vivo thermodynamics of the ED pathway in Zymomonas mobilis. Metab Eng 54:301–316. doi: 10.1016/j.ymben.2019.05.006. [DOI] [PubMed] [Google Scholar]
  • 19.Hermans MA, Neuss B, Sahm H. 1991. Content and composition of hopanoids in Zymomonas mobilis under various growth conditions. J Bacteriol 173:5592–5595. doi: 10.1128/jb.173.17.5592-5595.1991. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 20.Welander PV, Hunter RC, Zhang L, Sessions AL, Summons RE, Newman DK. 2009. Hopanoids play a role in membrane integrity and pH homeostasis in Rhodopseudomonas palustris TIE-1. J Bacteriol 191:6145–6156. doi: 10.1128/JB.00460-09. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 21.Schmerk CL, Bernards MA, Valvano MA. 2011. Hopanoid production is required for low-pH tolerance, antimicrobial resistance, and motility in Burkholderia cenocepacia. J Bacteriol 193:6712–6723. doi: 10.1128/JB.05979-11. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 22.Kulkarni G, Busset N, Molinaro A, Gargani D, Chaintreuil C, Silipo A, Giraud E, Newman DK. 2015. Specific hopanoid classes differentially affect free-living and symbiotic states of Bradyrhizobium diazoefficiens. mBio 6:e01251-15. doi: 10.1128/mBio.01251-15. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 23.Sáenz JP, Grosser D, Bradley AS, Lagny TJ, Lavrynenko O, Broda M, Simons K. 2015. Hopanoids as functional analogues of cholesterol in bacterial membranes. Proc Natl Acad Sci U S A 112:11971–11976. doi: 10.1073/pnas.1515607112. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 24.Wu C-H, Bialecka-Fornal M, Newman DK. 2015. Methylation at the C-2 position of hopanoids increases rigidity in native bacterial membranes. Elife 4:e05663. doi: 10.7554/eLife.05663. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 25.Silipo A, Vitiello G, Gully D, Sturiale L, Chaintreuil C, Fardoux J, Gargani D, Lee H-I, Kulkarni G, Busset N, Marchetti R, Palmigiano A, Moll H, Engel R, Lanzetta R, Paduano L, Parrilli M, Chang W-S, Holst O, Newman DK, Garozzo D, D’Errico G, Giraud E, Molinaro A. 2014. Covalently linked hopanoid-lipid A improves outer-membrane resistance of a Bradyrhizobium symbiont of legumes. Nat Commun 5:5106. doi: 10.1038/ncomms6106. [DOI] [PubMed] [Google Scholar]
  • 26.Flesch G, Rohmer M. 1987. Growth inhibition of hopanoid synthesizing bacteria by squalene cyclase inhibitors. Arch Microbiol 147:100–104. doi: 10.1007/BF00492912. [DOI] [Google Scholar]
  • 27.Horbach S, Neuss B, Sahm H. 1991. Effect of azasqualene on hopanoid biosynthesis and ethanol tolerance of Zymomonas mobilis. FEMS Microbiol Lett 79:347–350. doi: 10.1111/j.1574-6968.1991.tb04553.x. [DOI] [Google Scholar]
  • 28.Brenac L, Baidoo EEK, Keasling JD, Budin I. 2019. Distinct functional roles for hopanoid composition in the chemical tolerance of Zymomonas mobilis. Mol Microbiol 112:1564–1575. doi: 10.1111/mmi.14380. [DOI] [PubMed] [Google Scholar]
  • 29.Skerker JM, Leon D, Price MN, Mar JS, Tarjan DR, Wetmore KM, Deutschbauer AM, Baumohl JK, Bauer S, Ibáñez AB, Mitchell VD, Wu CH, Hu P, Hazen T, Arkin AP. 2013. Dissecting a complex chemical stress: chemogenomic profiling of plant hydrolysates. Mol Syst Biol 9:674. doi: 10.1038/msb.2013.30. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 30.Schmidt A, Bringer-Meyer S, Poralla K, Sahm H. 1986. Effect of alcohols and temperature on the hopanoid content of Zymomonas mobilis. Appl Microbiol Biotechnol 25:32–36. doi: 10.1007/BF00252509. [DOI] [Google Scholar]
  • 31.Qi LS, Larson MH, Gilbert LA, Doudna JA, Weissman JS, Arkin AP, Lim WA. 2013. Repurposing CRISPR as an RNA-guided platform for sequence-specific control of gene expression. Cell 152:1173–1183. doi: 10.1016/j.cell.2013.02.022. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 32.Bikard D, Jiang W, Samai P, Hochschild A, Zhang F, Marraffini LA. 2013. Programmable repression and activation of bacterial gene expression using an engineered CRISPR-Cas system. Nucleic Acids Res 41:7429–7437. doi: 10.1093/nar/gkt520. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 33.Rath D, Amlinger L, Hoekzema M, Devulapally PR, Lundgren M. 2015. Efficient programmable gene silencing by Cascade. Nucleic Acids Res 43:237–246. doi: 10.1093/nar/gku1257. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 34.Zheng Y, Han J, Wang B, Hu X, Li R, Shen W, Ma X, Ma L, Yi L, Yang S, Peng W. 2019. Characterization and repurposing of the endogenous Type I-F CRISPR–Cas system of Zymomonas mobilis for genome engineering. Nucleic Acids Res 47:11461–11475. doi: 10.1093/nar/gkz940. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 35.Peters JM, Colavin A, Shi H, Czarny TL, Larson MH, Wong S, Hawkins JS, Lu CHS, Koo B-M, Marta E, Shiver AL, Whitehead EH, Weissman JS, Brown ED, Qi LS, Huang KC, Gross CA. 2016. A comprehensive, CRISPR-based functional analysis of essential genes in bacteria. Cell 165:1493–1506. doi: 10.1016/j.cell.2016.05.003. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 36.Vigouroux A, Bikard D. 2020. CRISPR tools to control gene expression in bacteria. Microbiol Mol Biol Rev 84:e00077-19. doi: 10.1128/MMBR.00077-19. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 37.Wang T, Guan C, Guo J, Liu B, Wu Y, Xie Z, Zhang C, Xing X-H. 2018. Pooled CRISPR interference screening enables genome-scale functional genomics study in bacteria with superior performance. Nat Commun 9:2475. doi: 10.1038/s41467-018-04899-x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 38.Liu X, Gallay C, Kjos M, Domenech A, Slager J, van Kessel SP, Knoops K, Sorg RA, Zhang J, Veening J. 2017. High‐throughput CRISPRi phenotyping identifies new essential genes in Streptococcus pneumoniae. Mol Syst Biol 13:931. doi: 10.15252/msb.20167449. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 39.Müh U, Pannullo AG, Weiss DS, Ellermeier CD. 2019. A xylose-inducible expression system and a CRISPR interference plasmid for targeted knockdown of gene expression in Clostridioides difficile. J Bacteriol 201:e00711-18. doi: 10.1128/JB.00711-18. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 40.Rousset F, Cui L, Siouve E, Becavin C, Depardieu F, Bikard D. 2018. Genome-wide CRISPR-dCas9 screens in E. coli identify essential genes and phage host factors. PLoS Genet 14:e1007749. doi: 10.1371/journal.pgen.1007749. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 41.Peters JM, Koo B-M, Patino R, Heussler GE, Hearne CC, Qu J, Inclan YF, Hawkins JS, Lu CHS, Silvis MR, Harden MM, Osadnik H, Peters JE, Engel JN, Dutton RJ, Grossman AD, Gross CA, Rosenberg OS. 2019. Enabling genetic analysis of diverse bacteria with Mobile-CRISPRi. Nat Microbiol 4:244–250. doi: 10.1038/s41564-018-0327-z. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 42.Qu J, Prasad NK, Yu MA, Chen S, Lyden A, Herrera N, Silvis MR, Crawford E, Looney MR, Peters JM, Rosenberg OS. 2019. Modulating pathogenesis with Mobile-CRISPRi. J Bacteriol 201:e00304-19. doi: 10.1128/JB.00304-19. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 43.Roberts RJ, Vincze T, Posfai J, Macelis D. 2015. REBASE—a database for DNA restriction and modification: enzymes, genes and genomes. Nucleic Acids Res 43:D298–D299. doi: 10.1093/nar/gku1046. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 44.Yang Y, Shen W, Huang J, Li R, Xiao Y, Wei H, Chou Y-C, Zhang M, Himmel ME, Chen S, Yi L, Ma L, Yang S. 2019. Prediction and characterization of promoters and ribosomal binding sites of Zymomonas mobilis in system biology era. Biotechnol Biofuels 12:52. doi: 10.1186/s13068-019-1399-6. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 45.Pédelacq J-D, Cabantous S, Tran T, Terwilliger TC, Waldo GS. 2006. Engineering and characterization of a superfolder green fluorescent protein. Nat Biotechnol 24:79–88. doi: 10.1038/nbt1172. [DOI] [PubMed] [Google Scholar]
  • 46.Sadler JR, Sasmor H, Betz JL. 1983. A perfectly symmetric lac operator binds the lac repressor very tightly. Proc Natl Acad Sci U S A 80:6785–6789. doi: 10.1073/pnas.80.22.6785. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 47.Vera JM, Ghosh IN, Zhang Y, Hebert AS, Coon JJ, Landick R. 2020. Genome-scale transcription-translation mapping reveals features of Zymomonas mobilis transcription units and promoters. mSystems 5:e00250-20. doi: 10.1128/mSystems.00250-20. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 48.Hawley DK, McClure WR. 1983. Compilation and analysis of Escherichia coli promoter DNA sequences. Nucleic Acids Res 11:2237–2255. doi: 10.1093/nar/11.8.2237. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 49.Estrem ST, Gaal T, Ross W, Gourse RL. 1998. Identification of an UP element consensus sequence for bacterial promoters. Proc Natl Acad Sci U S A 95:9761–9766. doi: 10.1073/pnas.95.17.9761. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 50.Hawkins JS, Silvis MR, Koo B-M, Peters JM, Jost M, Hearne CC, Weissman JS, Todor H, Gross CA. 2019. Modulated efficacy CRISPRi reveals evolutionary conservation of essential gene expression-fitness relationships in bacteria. bioRxiv doi: 10.1101/805333. [DOI] [PMC free article] [PubMed]
  • 51.Jost M, Santos DA, Saunders RA, Horlbeck MA, Hawkins JS, Scaria SM, Norman TM, Hussmann JA, Liem CR, Gross CA, Weissman JS. 2020. Titrating gene expression using libraries of systematically attenuated CRISPR guide RNAs. Nat Biotechnol 38:355–364. doi: 10.1038/s41587-019-0387-5. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 52.Larson MH, Gilbert LA, Wang X, Lim WA, Weissman JS, Qi LS. 2013. CRISPR interference (CRISPRi) for sequence-specific control of gene expression. Nat Protoc 8:2180–2196. doi: 10.1038/nprot.2013.132. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 53.Gilbert LA, Horlbeck MA, Adamson B, Villalta JE, Chen Y, Whitehead EH, Guimaraes C, Panning B, Ploegh HL, Bassik MC, Qi LS, Kampmann M, Weissman JS. 2014. Genome-scale CRISPR-mediated control of gene repression and activation. Cell 159:647–661. doi: 10.1016/j.cell.2014.09.029. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 54.Wool IG. 1979. The structure and function of eukaryotic ribosomes. Annu Rev Biochem 48:719–754. doi: 10.1146/annurev.bi.48.070179.003443. [DOI] [PubMed] [Google Scholar]
  • 55.Wecker MS, Zall RR. 1987. Production of acetaldehyde by Zymomonas mobilis. Appl Environ Microbiol 53:2815–2820. doi: 10.1128/AEM.53.12.2815-2820.1987. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 56.Liu Y, Ghosh IN, Martien J, Zhang Y, Amador-Noguez D, Landick R. 2020. Regulated redirection of central carbon flux enhances anaerobic production of bioproducts in Zymomonas mobilis. Metab Eng 61:261–274. doi: 10.1016/j.ymben.2020.06.005. [DOI] [PubMed] [Google Scholar]
  • 57.Khandelwal R, Agrawal S, Singhi D, Srivastava P, Bisaria VS. 2018. Deletion of pyruvate decarboxylase gene in Zymomonas mobilis by recombineering through bacteriophage lambda red genes. J Microbiol Methods 151:111–117. doi: 10.1016/j.mimet.2018.06.008. [DOI] [PubMed] [Google Scholar]
  • 58.Seo J-S, Chong H-Y, Kim J-H, Kim J-Y. June 2009. Method for mass production of primary metabolites, strain for mass production of primary metabolites, and method for preparation thereof. US patent US20090162910A1.
  • 59.Bradley AS, Pearson A, Sáenz JP, Marx CJ. 2010. Adenosylhopane: the first intermediate in hopanoid side chain biosynthesis. Org Geochem 41:1075–1081. doi: 10.1016/j.orggeochem.2010.07.003. [DOI] [Google Scholar]
  • 60.Ong WK, Courtney DK, Pan S, Andrade RB, Kiley PJ, Pfleger BF, Reed JL. 2020. Model-driven analysis of mutant fitness experiments improves genome-scale metabolic models of Zymomonas mobilis ZM4. PLoS Comput Biol 16:e1008137. doi: 10.1371/journal.pcbi.1008137. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 61.Bringer S, Härtner T, Poralla K, Sahm H. 1985. Influence of ethanol on the hopanoid content and the fatty acid pattern in batch and continuous cultures of Zymomonas mobilis. Arch Microbiol 140:312–316. doi: 10.1007/BF00446969. [DOI] [Google Scholar]
  • 62.Baba T, Ara T, Hasegawa M, Takai Y, Okumura Y, Baba M, Datsenko KA, Tomita M, Wanner BL, Mori H. 2006. Construction of Escherichia coli K-12 in-frame, single-gene knockout mutants: the Keio collection. Mol Syst Biol 2:2006.0008. doi: 10.1038/msb4100050. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 63.Koo B-M, Kritikos G, Farelli JD, Todor H, Tong K, Kimsey H, Wapinski I, Galardini M, Cabal A, Peters JM, Hachmann A-B, Rudner DZ, Allen KN, Typas A, Gross CA. 2017. Construction and analysis of two genome-scale deletion libraries for Bacillus subtilis. Cell Syst 4:291–305.e7. doi: 10.1016/j.cels.2016.12.013. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 64.Calvo-Villamañán A, Ng JW, Planel R, Ménager H, Chen A, Cui L, Bikard D. 2020. On-target activity predictions enable improved CRISPR-dCas9 screens in bacteria. Nucleic Acids Res 48:e64. doi: 10.1093/nar/gkaa294. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 65.Datsenko KA, Wanner BL. 2000. One-step inactivation of chromosomal genes in Escherichia coli K-12 using PCR products. Proc Natl Acad Sci U S A 97:6640–6645. doi: 10.1073/pnas.120163297. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 66.Blodgett JAV, Thomas PM, Li G, Velasquez JE, van der Donk WA, Kelleher NL, Metcalf WW. 2007. Unusual transformations in the biosynthesis of the antibiotic phosphinothricin tripeptide. Nat Chem Biol 3:480–485. doi: 10.1038/nchembio.2007.9. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 67.Ghosh IN, Martien J, Hebert AS, Zhang Y, Coon JJ, Amador-Noguez D, Landick R. 2019. OptSSeq explores enzyme expression and function landscapes to maximize isobutanol production rate. Metab Eng 52:324–340. doi: 10.1016/j.ymben.2018.12.008. [DOI] [PubMed] [Google Scholar]
  • 68.Khan SR, Gaines J, Roop RM, Farrand SK. 2008. Broad-host-range expression vectors with tightly regulated promoters and their use to examine the influence of TraR and TraM expression on Ti plasmid quorum sensing. Appl Environ Microbiol 74:5053–5062. doi: 10.1128/AEM.01098-08. [DOI] [PMC free article] [PubMed] [Google Scholar]

Associated Data

This section collects any data citations, data availability statements, or supplementary materials included in this article.

Supplementary Materials

Supplemental file 1
AEM.01621-20-s0001.pdf (908.5KB, pdf)

Data Availability Statement

Plasmids and their sequences are available from Addgene (Addgene identification numbers 160073 to 160080).


Articles from Applied and Environmental Microbiology are provided here courtesy of American Society for Microbiology (ASM)

RESOURCES