Sunteți pe pagina 1din 39

Resource

A Comprehensive, CRISPR-based Functional


Analysis of Essential Genes in Bacteria
Graphical Abstract Authors
Jason M. Peters, Alexandre Colavin,
Handuo Shi, ..., Lei S. Qi,
Kerwyn Casey Huang, Carol A. Gross

Correspondence
stanley.qi@stanford.edu (L.S.Q.),
kchuang@stanford.edu (K.C.H.),
cgrossucsf@gmail.com (C.A.G.)

In Brief
A systematic analysis of all essential
genes in Bacillus subtilis using a CRISPR-
based knockdown approach established
the essential gene network, identified
modes of action for antibiotics, and
discerned fundamental underpinnings of
growth and morphological
characteristics of cells.

Highlights Accession Numbers


d CRISPRi knockdown of essential genes enables discovery of GSE74926
direct antibiotic targets

d An essential gene network reveals functional connections


between core processes

d A cell morphology screen identifies essential genes


intimately tied to cell shape

d A majority of essential genes show morphological defects as


a terminal phenotype

Peters et al., 2016, Cell 165, 1493–1506


June 2, 2016 Published by Elsevier Inc.
http://dx.doi.org/10.1016/j.cell.2016.05.003
Resource

A Comprehensive, CRISPR-based Functional Analysis


of Essential Genes in Bacteria
Jason M. Peters,1,15 Alexandre Colavin,2,15 Handuo Shi,3,15 Tomasz L. Czarny,4,11 Matthew H. Larson,5,6,7,8
Spencer Wong,5 John S. Hawkins,1,9 Candy H.S. Lu,1 Byoung-Mo Koo,1 Elizabeth Marta,1 Anthony L. Shiver,1,9
Evan H. Whitehead,5,7,10 Jonathan S. Weissman,5,6,7,8,10 Eric D. Brown,4,11 Lei S. Qi,3,12,13,* Kerwyn Casey Huang,2,3,14,*
and Carol A. Gross1,*
1Department of Microbiology and Immunology, University of California, San Francisco, San Francisco, CA 94158, USA
2Biophysics Program, Stanford University, Stanford, CA 94305, USA
3Department of Bioengineering, Stanford University, Stanford, CA 94305, USA
4Department of Biochemistry and Biomedical Sciences, McMaster University, Hamilton, ON L8S 3Z5, Canada
5Department of Cellular and Molecular Pharmacology, University of California, San Francisco, San Francisco, CA 94158, USA
6Howard Hughes Medical Institute, University of California, San Francisco, San Francisco, CA 94158, USA
7California Institute for Quantitative Biomedical Research, San Francisco, CA 94158, USA
8Center for RNA Systems Biology, University of California, Berkeley, Berkeley, CA 94720, USA
9Biophysics Graduate Program, University of California, San Francisco, San Francisco, CA 94158, USA
10UCSF Center for Systems and Synthetic Biology, University of California, San Francisco, San Francisco, CA 94158, USA
11Michael G. DeGroote Institute of Infectious Disease Research, McMaster University, Hamilton, ON L8N 3Z5, Canada
12Department of Chemical and Systems Biology, Stanford University, Stanford, CA 94305, USA
13ChEM-H, Stanford University, Stanford, CA 94305, USA
14School of Medicine, Department of Microbiology and Immunology, Stanford University, Stanford, CA 94305, USA
15Co-first author

*Correspondence: stanley.qi@stanford.edu (L.S.Q.), kchuang@stanford.edu (K.C.H.), cgrossucsf@gmail.com (C.A.G.)


http://dx.doi.org/10.1016/j.cell.2016.05.003

SUMMARY facilitating drug development. Yet, few approaches can assess


essential gene function in vivo to elucidate their connections.
Essential gene functions underpin the core reactions Neither gene-deletion libraries (Baba et al., 2006; Winzeler
required for cell viability, but their contributions et al., 1999) nor saturating transposon mutagenesis (Goodman
and relationships are poorly studied in vivo. Using et al., 2009; van Opijnen et al., 2009) can be used to study
CRISPR interference, we created knockdowns of essential genes, as cells cannot survive without their functions
every essential gene in Bacillus subtilis and probed (Christen et al., 2011). Several high-throughput approaches
have been used to identify or perturb essential genes in eukary-
their phenotypes. Our high-confidence essential
otes, including destabilizing the 30 UTR of mRNAs (DaMP alleles;
gene network, established using chemical genomics,
Breslow et al., 2008), CRISPR (clustered regularly interspaced
showed extensive interconnections among distantly short palindromic repeats)/Cas9 gene editing (Blomen et al.,
related processes and identified modes of action for 2015; Wang et al., 2015), and CRISPR/dCas9 transcriptional
uncharacterized antibiotics. Importantly, mild knock- regulation technologies (Gilbert et al., 2014). The only study
down of essential gene functions significantly reduced screening essential genes in bacteria used antisense RNA
stationary-phase survival without affecting maximal knockdowns to screen for antibiotic sensitivities (Xu et al.,
growth rate, suggesting that essential protein levels 2010), a method of limited utility due to variable efficacy (Forsyth
are set to maximize outgrowth from stationary phase. et al., 2002).
Finally, high-throughput microscopy indicated that Here, we establish a CRISPR interference (CRISPRi) frame-
cell morphology is relatively insensitive to mild knock- work for systematic phenotypic analysis of essential genes
in bacteria. CRISPRi uses a nuclease-deactivated variant of
down but profoundly affected by depletion of gene
Streptococcus pyogenes Cas9 (dCas9) paired with a single
function, revealing intimate connections between cell
guide RNA (sgRNA) to sterically hinder transcription at the
growth and shape. Our results provide a framework sgRNA base-pairing genomic locus (Qi et al., 2013) and is a
for systematic investigation of essential gene func- specific and efficient approach for knockdown, with demon-
tions in vivo broadly applicable to diverse microorgan- strated applicability in bacteria. We generated a comprehen-
isms and amenable to comparative analysis. sive essential gene-knockdown library in the Gram-positive
model bacterium Bacillus subtilis and used the library to enable
INTRODUCTION drug target discovery, establish a functional network of essen-
tial gene processes, characterize how cell morphology and
Essential gene functions underpin core cellular processes. Inter- growth rate respond to reductions in essential gene expres-
rogating the relationships among essential gene functions is crit- sion, and dissect essentiality in a highly redundant genetic
ical for understanding how bacterial growth is controlled and for pathway. Our study provides a framework for comprehensive,

Cell 165, 1493–1506, June 2, 2016 Published by Elsevier Inc. 1493


CRISPRi for targeted gene repression in B. subtilis Figure 1. B. subtilis CRISPRi Is Efficient,
A
Titratable, and Specific
Inducible promoter
(A) B. subtilis xylose-inducible dCas9 is directed to
Pxyl specific DNA targets by constitutively expressed
dCas9
sgRNAs, where it represses transcription. dcas9
dcas9
was stably integrated into the lacA locus, and
lacA´ ´lacA
* sgRNAs were integrated into amyE or thrC.
Transcriptional
* (B) Flow cytometry of cells that constitutively ex-
interference
press rfp and an rfp-targeted sgRNA, in which
Pveg dcas9 was induced by adding the specified con-
sgRNA * centration of xylose. Median RFP levels are relative
RNAP
sgRNA to the no-sgRNA control (gray dashed line), and
amyE´ ´amyE error bars are ± 1 SD.
Base-pairing dCas9 (C) RNA sequencing (RNA-seq) of cells maximally
region handle
induced for dcas9 (1% xylose) and constitutively
expressing rfp and an rfp-targeted sgRNA versus
B C sgRNA+ (normalized reads/gene) cells without an sgRNA. Reads per gene were
1 normalized by reads per kilobase per million reads.
~3X

~150X
Relative RFP expression

The dashed line is y = x.


100 (D) NET-seq of cells maximally induced for dcas9
(1% xylose) and constitutively expressing rfp and
0.1 an rfp-targeted sgRNA versus cells without an
10 sgRNA. The y axis is broken to accommodate the
wide range of reads. The dashed line corresponds
rfp
1 to the upstream boundary of the PAM sequence.
We suggest that RNA 30 end peaks upstream of the
0.01 dCas9 block are the result of RNA polymerase
0.1 R = 0.99 queuing.
0.001 0.01 0.1 1 0.1 1 10 100 (E) Flow cytometry of cells maximally induced
% xylose sgRNA- (normalized reads/gene) for dcas9 (1% xylose) and constitutively express-
ing an rfp-gfp operon and various rfp-targeting
sgRNAs. RFP and GFP levels are relative to the no-
D 94 122 E sgRNA control. The dashed line is y = x. Note that
PAM rfp-gfp operon
sgRNA binding site 1 in this group of sgRNAs, only template strand-
Normalized RNA 3´-ends (×1000)

rfp rfp gfp targeting sgRNAs exhibited low efficacy (violet).


Relative GFP expression

3
template See also Figure S1 and Table S1.
sgRNA-
2
non-template
1 0.1
0
100
sgRNA+
50 (IPTG)-inducible promoter (Figure S1B),
20 we exclusively used Pxyl throughout to
10 0.01
2
consistently sensitize the library with slight
1
R = 0.99
knockdown and avoid inconsistencies
0 resulting from slight variations in IPTG
0 100 200 300 400 500 600 0.01 0.1 1
Distance from rfp start codon (bp) Relative RFP expression induction.
Our Pxyl-based CRISPRi system is
titratable, with unimodal rfp repression
high-throughput analysis of essential gene functions applicable at the single-cell level across sub-saturating inducer concen-
to diverse bacteria. trations (Figures 1B and S1C), and high specificity, repressing
only rfp expression at saturating inducer concentrations (Fig-
RESULTS ure 1C). We used NET-seq to identify the genomic positions
of transcribing RNA polymerase (Larson et al., 2014), showing
CRISPRi Is Effective, Specific, and Titratable in that CRISPRi sterically blocks transcription in B. subtilis (Fig-
B. subtilis ures 1D and S1D), as in Escherichia coli (Qi et al., 2013).
We established a CRISPRi system in B. subtilis, consisting of CRISPRi is polar (Peters et al., 2015), with all downstream
S. pyogenes dcas9 driven by a xylose-inducible promoter (Pxyl) genes in an operon showing equivalent knockdown (Fig-
and sgRNAs expressed from a strong, constitutive promoter ure 1E), and also reduces expression of upstream genes in
(Pveg), both transferred to the chromosome via integrating plas- the operon (Figure S1E). Thus, CRISPRi technology is suitable
mids (Figure 1A). This system is very efficient, exhibiting 3-fold for examining gene function at the operon level.
repression of red fluorescent protein (RFP) without induction
and 150-fold repression with full dcas9 induction (Figures 1B A CRISPRi Knockdown Library of Essential Genes
and S1A). Although we also report a system with no basal repres- We constructed an arrayed library of B. subtilis strains expressing
sion based on a weak isopropyl b-D-1-thiogalactopyranoside computationally optimized sgRNAs (Supplemental Experimental

1494 Cell 165, 1493–1506, June 2, 2016


A B sgRNAuppS + 0.05% xyl Figure 2. CRISPRi Knockdowns of Essen-
O
sgRNAuppS tial Genes Enable Discovery of Direct Antibi-
# of CRISPRi knockdown strains

O 140 wt otic Targets


N N
30 Br uppS (A) Relative fitness of CRISPRi essential gene
O 120
MAC-0170636 knockdown strains (n = 289) with basal dcas9
100 expression (no xylose induction) grown on plates

Growth (%)
20 containing MAC-0170636, as determined by the
80
ratio of normalized colony sizes on LB plates +
60 DMSO versus LB + MAC-0170636.
uppS (B) Minimal inhibitory concentration (MIC) assay
10 40 for strains over- or underexpressing uppS grown in
liquid medium containing MAC-0170636. The
20
MICs of these strains were (in mg/ml) sgRNAuppS +
0 0 0.05% xyl (0.012), sgRNAuppS (0.195), WT
0 0.5 1 1.5 0.001 0.01 0.1 1 10 (0.78), and uppS overexpression (3.125). The
Sensitive Relative fitness Resistant MAC-0170636 (μg/mL) plotted values are the means of at least three
measurements, and error bars are ± 1 SD.
(C) Concentration-dependent inhibition of purified
C D B. subtilis B. subtilis UppS by MAC-0170636. Each point is
B. subtilis with uppSBs an individual measurement.
120 S. aureus
140 (D) MIC assay for strains expressing B. subtilis uppS
B. subtilis with uppSSa
(BsDuppS/amyE::Pspank-uppSBs) or S. aureus uppS
120
100 (BsDuppS/amyE::Pspank-uppSSa) grown in liquid
UppS activity (%)

100 LB + MAC-0170636.
Growth (%)

80
See also Figure S2 and Tables S1 and S2.
80

60 60

40
40 IC50 = 0.79 µM
2011). Most knockdowns of antibiotic
20
targets were hypersensitized to their
20 0 cognate drug (e.g., dfrA/folate biosyn-
-4 -3 -2 -1 1 2
10 10 10 10 1 10 10 0.01 0.1 1 10
thesis to trimethoprim (Myoda et al.,
MAC-0170636 (μM) MAC-0170636 (μg/mL) 1984) and fabF/fatty acid metabolism
to cerulenin (Moche et al., 1999) (Table
S2). Although fabI, the target of triclosan
Procedures) targeting the 289 known or proposed essential (Schujman et al., 2001), is not essential due to a gene duplication
genes (Table S1). The sgRNAs targeted unique DNA sequences (Thomaides et al., 2007), another knockdown in the pathway
at the 50 ends of genes, where CRISPRi is most effective (Qi (fabG) was sensitized (Table S2). We conclude that our CRISPRi
et al., 2013). Nearly all sgRNAs (94%) targeting bona fide essen- platform effectively identifies known drug-gene interactions.
tial genes (258 genes total; Supplemental Experimental Proce- We tested our essential knockdown library as a platform
dures) (B.M.K., J.M.P., and C.A.G., unpublished data) decreased for drug target discovery by screening against MAC-0170636,
colony size on agar plates with xylose (R25% reduction in area an antibiotic that upregulates the cell-wall-damage-responsive
compared to the control; Table S1; Supplemental Experimental promoter PywaC (Czarny et al., 2014) by an unknown mechanism.
Procedures). Control cells expressing only dcas9 or dcas9 and Undecaprenyl pyrophosphate synthetase (uppS) was the most
an sgRNA targeting an innocuous gene (rfp) had no growth de- sensitized knockdown (Figure 2A; Table S2), and we confirmed
fects (Figure S1F). We conclude that a single sgRNA is sufficient its sensitivity in liquid (Figure 2B). Conversely, uppS overexpres-
for effective knockdown of essential genes, simplifying CRISPRi sion increased MAC-0170636 resistance relative to wild-type
library design in bacteria. (WT) cells (Figure 2B). Purified B. subtilis UppS activity was in-
hibited by MAC-0170636 with an IC50 of 0.79 mM (Figure 2C),
CRISPRi-Based Essential Gene Phenotyping and Drug indicating that UppS is the direct target of MAC-0170636.
Target Discovery UppS purified from another Firmicute, Staphylococcus aureus,
The 3-fold repression of our knockdown library without induc- was completely resistant to MAC-017063 (Figure S2E), as was
tion (basal repression; Figures 1B and S1A) sensitized strains to S. aureus itself. Likewise, B. subtilis expressing only S. aureus
various chemicals. This enabled us to define essential gene phe- uppS was resistant (Figure 2D). These results highlight the utility
notypes via chemical-genomic analysis by measuring colony of our knockdown library for identifying direct targets of unchar-
size against 35 unique compounds (Supplemental Experimental acterized compounds; CRISPRi portability suggests its utility in
Procedures). We achieved high reproducibility, as measured by future organism-specific drug-discovery efforts.
correlated colony sizes (R = 0.89; Figure S2A). We converted
colony sizes to chemical-gene scores (Figures S2A–S2D; Table Functional Analysis of the Essential Gene Network
S2; Nichols et al., 2011) and identified significant chemical- Highly correlated responses of gene knockdowns across chemical
gene phenotypes (false discovery rate %5%; Nichols et al., conditions (phenotypic signatures) indicate functional connections

Cell 165, 1493–1506, June 2, 2016 1495


A B. subtilis essential gene network B

phenotype correlation
0.57 0.89
Cell division

Cell wall biosynthesis

C
1

True positive rate (TPR)


0.8 TPR/FPR ≈ 19
0.3
0.6 FPR

ylaN 0.4 0.15

TPR
phenotype correlation 0.2
cell wall
0.57 0.94 0
cell division 0 0.05
RNA polymerase between genes/operons 0
isoprenoid biosynthesis (MEP) within operons 0 0.2 0.4 0.6 0.8 1
iron-sulfur cluster biosynthesis False positive rate (FPR)
unknown

D E # of inter-process connections
STRING database essential network
% of essential network connections

0 1 2 3 4 5+ 0 1 2 3 4 5+
70 present in STRING DNA condensation/segregation: 1
60 absent from STRING DNA replication: 2
RNases: 3
50 aminoacyl-tRNA synthetases: 4
biosynthesis of fatty acids: 5
40 biosynthesis of isoprenoids: 6
biosynthesis of peptidoglycan: 7
30 cell division: 8
protein secretion: 9
20 ribosomal proteins: 10
ribosome assembly: 11
10 tRNA modification and maturation: 12
translation factors: 13
0
0 1 2 3 4 1 2 3 4 5 6 7 8 9 10 11 12 13
Annotation distance

Figure 3. An Essential Gene Network Reveals Numerous Intra- and Inter-process Connections
(A) Essential gene network based on correlations between chemical-gene phenotypes. Edge thickness is proportional to the extent of correlation. See also
Figure S3 for network gene names.
(B) Intra- and inter-process connections between cell division and cell-wall biosynthesis genes. Genes outside the main network or genes lacking intra-process
connections were excluded.
(C) ROC curve comparing connections between essential operons in our network to the STRING database. True-positive connections are those present in the
high-confidence set of interactions from STRING, and false-positive connections are absent from STRING (these connections may be either truly false or novel).
(D) Annotation distance for network connections between essential operons present or absent from the STRING database. Genes with an annotation distance of
zero are from the same functional group, while genes with an annotation distance of four are unconnected by annotation.
(E) Functional annotations for intra-process connections between essential operons present in the STRING database or present in our essential network.
See also Figures S2, S3, and S4 and Tables S1, S2, and S3.

(Nichols et al., 2011). We established a network of gene-gene 3A, S2B–S2D, and S3; Table S3; Supplemental Experimental Pro-
connections using statistically significant correlations among the cedures). The network was rich in known biological connections
phenotypic signatures of our essential gene knockdowns, based among genes in related processes, such as cell-wall biosynthesis
on direct or indirect effects of drug-gene interactions (Figures and cell division (Figure 3B).

1496 Cell 165, 1493–1506, June 2, 2016


We quantitatively validated the network with a receiver oper- tions may be involved in failsafe mechanisms that link division
ating characteristic (ROC) curve. Because CRISPRi exhibits po- and DNA replication (Arjes et al., 2014).
larity, this analysis was based on operons containing essential We explored the unexpected connection between transcrip-
genes (‘‘essential operons,’’ n = 203), rather than on the individ- tion (rpoB) and cell division (ftsL), which is based on shared
ual genes themselves (Table S3; Supplemental Experimental sensitivities to DNA intercalators and cell-wall antibiotics (Fig-
Procedures). We compared high-confidence gene-gene con- ure S2D). Using RNaseq, we found that basal knockdown of
nections in the STRING database (http://string-db.org/; Szklarc- rpoB reduced the rpoB-rpoC mRNA level by two-fold. Among
zyk et al., 2015) with correlations between essential operons the few other substantial expression changes (Figure S4D), we
in our chemical-genomics dataset that met our correlation found several downregulated envelope genes including manA
threshold (R = 0.572; Figure 3C; Supplemental Experimental (4-fold), which is involved in cell-wall integrity (Elbaz and Ben-
Procedures). Our network showed excellent agreement with Yehuda, 2010), and sigW (1.5-fold to 2-fold), a master regulator
STRING, with a 20-fold ratio of true- to false-positive rates (Fig- of envelope stress (Cao et al., 2002) (Table S3). Selective reduc-
ure 3C, inset). tion of several envelope functions due to rpoB knockdown could
We investigated the relationship between network connec- result in cell-wall defects that mimic those caused by knock-
tions and gene function by assigning essential operons to func- down of late-acting cell-division genes.
tional groups using the SubtiWiki online resource (http:// We also identified novel connections to essential genes of
subtiwiki.uni-goettingen.de/; Michna et al., 2014). Of the unknown function (Figures 3A, S3, and S4E). For example, resis-
61 non-overlapping functional groups, the 35 with two or more tance to cell-wall-targeting antibiotics drove strong correlations
essential operons were assessed for connectivity within the among the largely uncharacterized gene ylaN (Figure S2F) (Hunt
group (intra-connectivity). Mean intra-connectivity was 2.4, et al., 2006; Xu et al., 2007), iron-sulfur cluster biogenesis, and
compared with a background mean connectivity of 0.3 isoprenoid biosynthesis, the latter of which depends on the
(p < 105), with ten groups showing connectivity among R3 iron-sulfur cluster enzyme IspH (Gräwert et al., 2004; Wolff
essential operons (Table S3). Indeed, the ROC analysis showed et al., 2003). These results suggest that defects in cellular iron
strong specificity and sensitivity for recapitulating existing intra- homeostasis underlie the connections to ylaN. Indeed, in follow
process connections in STRING (Figure S4A), with excellent up experiments, we determined that ylaN is non-essential with
recovery of connections within functional groups (Figure S4B). added iron(III) (Figure S4F), further highlighting the ability of our
Connections within the peptidoglycan (PG) cell-wall biosynthesis unbiased approach to identify novel connections among essen-
(18) and DNA replication (9) functional groups were most dense, tial processes.
likely reflecting convergence of PG precursor biosynthesis
and localized cell-wall assembly (Turner et al., 2014) and re- Growth Characteristics of the Essential Gene
plisome protein-protein interactions (Sanders et al., 2010), Knockdown Library
respectively. We measured apparent lag, maximum growth rate, and satu-
Next, we compared connectivity between functional groups rating density of growth curves for all basal knockdowns in
(inter-connectivity) in our network with those in STRING, finding lysogeny broth (LB) medium (Figures 4A and 4B; Supplemental
common connections and important distinctions (Figures 3D Experimental Procedures). Almost all knockdowns (80%) had
and 3E). First, the balance of intra- and inter-process connec- a maximal growth rate equivalent to the control (Figure 4B, inset);
tions differed; 46.2% of connections in STRING were inter-pro- ribosomal proteins were the only functional category with slower
cess versus 59.0% in our network (p < 104). Second, STRING growth rates relative to other knockdowns (p = 0.0008, t test).
inter-process connections were biased toward extensively stud- However, most knockdowns (95%) had longer apparent
ied processes, e.g., 39% of STRING, but only 4.3% of our inter- lag times (Figure 4A, inset), and cell-wall synthesis genes
process connections were between ribosomal proteins and were enriched for the longest lags (p = 0.0013, Mann-Whitney
translation factors. Moreover, 84% (113/134) of connections U test; Table S4). CRISPRi repression was maintained in station-
unique to our network were between processes (Table S3), high- ary phase (Figure S4G).
lighting the ability of our open-ended approach to detect such We determined whether longer apparent lag times reflected
connections. Finally, our network revealed many connections strain-specific differences or decreased viability. Both the frac-
between operons in distant functional groups. Using the hierar- tion of growing cells on agarose pads (Figure 4C) and plate
chical annotation levels from SubtiWiki (Michna et al., 2014) as viability measurements (Figure 4D) of strains spanning the
an approximation of ‘‘annotation distance,’’ we found that novel apparent lag times negatively correlated with apparent lag.
connections in our network were skewed toward processes Apparent lag times also tracked with expectations from the num-
furthest apart in annotation, whereas those in STRING were pre- ber of live cells present. For example, using the 45-min doubling
dominantly between related processes (Figure 3D). Some distant time in batch culture (Figure 4B, inset), the 8-fold reduction
connections were intuitive, such as DNA replication (holB) and in mraY knockdown viable cells (Figures 4C and 4D) requires
folate biosynthesis (folC and the sul-folB-folK operon), likely re- 3 additional divisions, accounting for the 2.1 hr increase in
flecting the folate requirement for deoxythymidine triphosphate apparent lag. Moreover, viable mraY knockdown cells and the
(dTTP) production (Hardy et al., 1987). Other connections were control had equivalent growth rates after transfer to an agarose
unexpected, such as those between peptidoglycan (PG) biosyn- pad with fresh LB (Figure 4C, inset). Together, these surprising
thesis/cell division and DNA replication/modification (e.g., ftsL results suggest that slight reductions in essential gene prod-
and dnaX, murC and gyrB, ddl-murF and ydiO); these connec- ucts affect outgrowth from stationary phase, creating a mixed

Cell 165, 1493–1506, June 2, 2016 1497


A B

Apparent lag time (h)


p < 10-16

Max. growth rate (h-1)


p = 0.15
4
control 1
1 1
3 essential genes
0.9

2
0.8

es
l
ro

en
nt

es
l
ro
OD600

OD600
co

lg

en
nt
ia

0.1

co

lg
0.1
nt

ia
se

nt
es

se
es
control
essential genes

0.01 0.01
0 1 2 3 4 5 6 7 -5 0 5
Time (h) Relative time to end of apparent lag (h)

C 0.04 D
Growth rate (min-1)

0.9 control
alrA 0.03 ×1010
plsX 5 plsX yneS
0.8
Proportion of growing cells

rpsJ hemY 0.02


yneS mraY
0.7 tyrS folK 0.01 4 control R = -0.73
prsA44 control
alrA folK p = 4 × 10-4
CFU/mL/OD600

0.6 rnpA 0 rpsJ


prsA44
0 20 40 60 80 100 120 3
0.5 yumC
rbgA Time on agar pad (min)
mreD hemY leuS rbgA
leuS iscU
0.4 2
tyrS yumC
rnpA acpP
0.3 R = -0.96 iscU hemA
p = 4 × 10-11 1
mreD
0.2
hemA acpP
mraY mraY
0.1 0
2 2.5 3 3.5 4 4.5 2 2.5 3 3.5 4 4.5
Apparent lag time (h) Apparent lag time (h)

Figure 4. High-Resolution Growth Profiles of Essential Gene Knockdowns Reveal Widespread Defects in Stationary-Phase Survival
(A) Microplate reader growth curves of essential gene knockdown library strains. We grew cells for 18 hr in LB before back-diluting into fresh LB (t = 0). The shaded
area is the mean ± SD for the no-sgRNA control, n = 23.
(B) Growth curves from (A) with the apparent lag time of each strain shifted to zero. The shaded area is the mean ± SD for the no-sgRNA control, n = 23.
(C) Data extracted from single-cell, time-lapse microscopy of selected essential gene knockdown strains grown on agarose pads with fresh LB after liquid growth
in LB for 18 hr. Although fewer mraY knockdown cells grow on LB pads after 18 hr (11% versus 87% for the no-sgRNA control), the growth rate of elongating cells
quantitatively matches that of the control (inset).
(D) Viable cell plating of selected essential gene knockdown strains grown in LB for 18 hr. CFU, colony-forming units.
See also Figure S4 and Tables S1 and S4.

population in which some cells exhibit WT outgrowth and others slower growth rate at the time of imaging, similar to previous re-
are either dead or non-growing. ports linking cell size and steady-state growth rate (Schaechter
et al., 1958) and consistent with known shortening in the station-
Basal Knockdown of Essential Genes Results in ary phase (Overkamp et al., 2015).
Changes in Cellular Dimensions that Reveal Shape Next, we determined morphological parameters of the essen-
Actuators and Modulators tial gene basal knockdowns to pinpoint proteins for which a
Our library offers an unprecedented opportunity to examine small expression change results in altered shape, using high-
whether basal knockdown of essential genes affects cell throughput imaging at 3.5 hr after dilution (Supplemental
morphology. To provide baseline values, we quantified the di- Experimental Procedures). Because of variable stationary-phase
mensions of WT cells. Cells were imaged 3.5 hr after different de- outgrowth (Figures 4C and 4D), culture OD widely varied at the
grees of dilution into fresh medium (Supplemental Experimental time of imaging. We found a general relationship between growth
Procedures) and therefore had different culture optical densities rate and OD for virtually all knockdowns (Figure S4H), indicating
(OD). We found that median cell length and width systematically a large range of instantaneous growth rates at imaging. Nonethe-
varied with the extent of dilution (Figure 5A), likely reflecting less, the length and width of the knockdowns were highly corre-
growth phase differences as cells cycle through nutrients in lated (Figure 5B), similar to WT cells across dilutions (Figure 5A).
rich medium (LB). Cells diluted less were smaller, reflecting their The fact that cells reach their maximum growth rate at a fixed OD

1498 Cell 165, 1493–1506, June 2, 2016


A B 13
8 ftsL

number of cells
sgRNA- 1:1280 nrdE
12 >400
7.5
1:640 11
1:320 mbl
7 10 <50 dnaX dnaA

Median length (μm)


Median length (μm)

1:160

9 murD
1:80
6.5
8 nrdF

6
7
1:40

5.5 6 tufA

5 mreB murE
5
1:20 4 cell envelope genes
murC replication or translation outliers
4.5 3
0.98 0.99 1 1.01 1.02 1.03 1.04 1.05 0.9 1.0 1.1 1.2
Median width (μm) Median width (μm)
C D 11.4 -

sgRNA- OD ~ 0.3
Median length (μm)
8 murE
11.0
OD ~ 0.3 sgRNA- sgRNA- murB
10.6
murC
7 murF
murAA
10.2
6
mbl murD
mreB 9.8
Fraction of cells

5
murG
9.4
1.02 1.04 1.06 1.08 1.1 1.12 1.14 1.16
4
Median width (μm)
E
mbl 1.12
3
Median width (μm)

murB
2
1.08 sgRNA-
5 μm
1.04
1
mreB 1
0
0.8 1 1.2 1.4 1.6
10-7 10-6 10-5 10-4 10-3 10-2
Width (μm)
% xylose

Figure 5. Partial Knockdown of Essential Genes Identifies Potential Morphological Regulators, with Envelope Gene Knockdown Leading to
Changes in Cell Width
(A) Cell volume of the no-sgRNA control substantially varies after 3.5 hr of growth across dilutions of the same overnight culture into fresh LB. Cell length and width
were highly correlated (Pearson’s R = 0.96, p < 0.001).
(B) Median cell length and width of essential knockdown strains are highly correlated after 3.5 hr of growth (Pearson’s R = 0.66, p < 1038). ‘‘Cell envelope’’ is the
only functional category enriched in the outliers to the best-fit line (black) between length and width. Selected non-cell envelope outliers in DNA replication (blue)
and translation (orange) are also shown.
(C) Distribution of cell widths of mbl and mreB knockdown strains at OD600 0.3 reveals subtle distinctions between the two actin homologs, with mbl cells wider
than no-sgRNA control cells and mreB cells adopting a broader range of widths. n > 2,000 cells for each histogram.
(D) Knockdown strains in the mur pathway have similar median cell length at OD600 0.3, while cell width varies over a wide range. Bars represent SEM.
(E) Cell width of the murB knockdown strain monotonically increases with the degree of dcas9 induction, in contrast to the no-sgRNA control.
See also Figure S5 and Tables S1 and S5.

(Figure S4H), rather than at a fixed number of doublings after p < 0.0005 by bootstrapping; Figure 5B; Supplemental Experi-
growth resumption, underscores the importance of cell density mental Procedures), as expected because they are largely PG
and the extracellular milieu in growth rate control. synthesis related and are therefore actuators of cell shape. To
Basal knockdowns that deviated from the length/width trend- validate our identification of outliers and ensure that cell chaining
line (Figure 5B; Table S5) could represent proteins involved in was rare, we used the membrane stain, FM4-64, on a subset of
either actuating or regulating growth of the shape-determining strains with different median cell lengths, finding that length
cell wall. Only cell-envelope genes exhibited significant enrich- measurements from phase images and peripheral fluorescence
ment of outliers (average median length deviation = 0.99 mm, were highly correlated (Figure S5A).

Cell 165, 1493–1506, June 2, 2016 1499


A
all strains (289) no defect bulging severe bending
morphological defects

69 11% 11%
10 μm
32%

growth defects
36%
5%
39 135 46 86%
39% 63%

i ii 14% iii iv 2%

growth and lysis growth halt filamentation overdivision


5% 11%
17%
11%
30%

52% 100%
83% 50%

29%
v vi 15% vii viii

unknown function metabolism central dogma envelope/division

B terminal phenotype similarity C


valS xylose titration
.: 1
b.s
c an on: 2 10 μm
gly ivisi : 3
do
pti cell d ation 4
pe ns n:
de tio
c on plica s: 5
A re se
DN NA RNa s: 6
D t a se 7
the n: LB 0.005% xylose 0.05% xylose 1% xylose
yn retio 8
N A s sec ins:
tR tein rote y: 9
yl- p l WT serine-hydroxamate titration
ac pro mal emb 10
i no so ass ion:
am o
rib me urat : 11
o at
os ors
rib on/m fact .: 12
i s
a t io id b. : 13
n
ific lat
m od rans ty ac b.s. 1 2 3 4 5 6 7 8 9 10 11 12 13
t fa t id
NA no
tR
o pre
is
less similar more similar LB 0.6 mg/mL 1 mg/mL 5 mg/mL

D E 0 min 20 min 40 min 60 min 80 min 100 min


tagD 5 μm
bulging
0.04
Growth rate (min-1)

pbpB
0.03
filamentation

0.02 accD
bending
0.01
murD
growth
ol

cD

aA

rD
rS

pB
E
l
X
mb

and Lysis
tag
ntr

rps
pls

se

mu
ac

mn

pb
co

Figure 6. Essential Gene Depletion Reveals a Diversity of Terminal Phenotypes


(A) (i) Area-proportional graph of the fraction of essential gene knockdowns that give rise to morphological and growth terminal phenotypes. (ii–viii) Single-cell
imaging of common terminal phenotypes of essential-gene knockdowns, with bar graphs depicting the broad functional categories underlying each terminal
phenotype.
(legend continued on next page)

1500 Cell 165, 1493–1506, June 2, 2016


Several outliers in other functional groups were intriguing, as We quantified terminal morphologies by manually classi-
they could identify cell-shape regulators. Basal knockdown of fying phenotypes with a 6D phenotype vector, including
tufA, encoding the translation elongation factor Tu, resulted in lysis, bulging, bending, filamentation, overdivision (shorter
cells that were substantially shorter than expected (Figure 5B; cells), and extent of growth. We assessed whether functionally
Table S5); Tu interacts with the bacterial cytoskeleton protein related genes mapped to particular phenotypes. Many cell-
and rod-shape determinant MreB (Defeu Soufo et al., 2010). envelope genes bulged (e.g., tagA/B/F/G/O [teichoic acid],
Knockdowns of several DNA replication genes (dnaX, dnaA, mreC [cell shape/PG biosynthesis]) (Figure 6Aiii), while many
and nrdE/F) resulted in longer cells (Figure 5B; Table S5). More- ribosomal genes displayed severe cell bending (Figure 6Aiv).
over, dnaX and ftsL, both of which are large mophological out- However, in some cases genes in different processes had
liers, were also significantly correlated in our chemical screen, similar terminal phenotypes, such as growth halting (Fig-
suggesting a functional connection between replication and ure 6Avi) and bending (Figure 6Aiv). Several metabolic genes
division, corroborated by multiple independent data types. (e.g., metK and dxs) exhibited bulging (Table S5), consistent
As cell size is dependent on the degree of dilution prior to re- with network connections between these processes and
growth (Figure 5A), direct comparisons of cellular dimensions cell-envelope synthesis/division (Figure 3A). Finally, some
among different strains require cells to be at the same OD. Large strains displayed both bulging and lysis (Table S5), indicating
variation in lag times prevented us from measuring the entire li- that depletions in one process can lead to more than one ter-
brary by this method. Instead, we examined two actin homologs minal phenotype.
(mreB and mbl) involved in coordinating PG synthesis (Scheffers Phenotypic variability within functional groups and common
and Pinho, 2005) and the mur genes responsible for PG precur- phenotypes of different groups made it challenging to map
sor synthesis at OD 0.3 ± 0.05. Both actin homologs were wider phenotype to process with simple visual comparison. Instead,
than WT, on average, with mreB exhibiting larger standard devi- we evaluated the similarities between terminal phenotype vec-
ation, suggesting that cell width is particularly sensitive to tors for all SubtiWiki annotation pairs (Figure 6B; Supplemental
MreB levels (Figure 5C). The median cell lengths of mur strains Experimental Procedures), a process conceptually similar to
were similar to WT (Figure 5D), but widths (population median the one we used to identify inter-process connections in the
between 1.04–1.14 mm) were larger than no-sgRNA cells chemical-genomics dataset. Our similarity matrix recapitulated
(1.02 mm), validating their classification as outliers (Figure 5D). known connections (e.g., PG synthesis/cell division), while
The murB cell width monotonically increased with dcas9 induc- providing support for novel network connections (e.g., fatty
tion (Figure 5E), indicating that cell width is responsive to the acid metabolism with both isoprenoid biosynthesis and several
cellular levels of PG precursors. central-dogma annotations; Figures 3A, 3B, and 6B). Terminal-
In summary, our systematic screen for the essential gene phenotype links are consistent with and complement the high
products most intimately tied to cell-shape identified genes level of inter-process connectivity of the essential gene network
with known ties to morphology (e.g., mreB/mbl) and cell-wall (Figure 3E).
synthesis, revealed a quantitative relationship between PG pre- Variable terminal phenotypes within a group might reflect the
cursor gene expression and cell width, and uncovered potential rate of gene-product depletion. We tested this hypothesis by
new regulators of cell morphology. inhibiting tRNA charging (Figure 6C), either by valS (ValS amino-
acyl-tRNA synthetase) knockdown or by addition of serine
Single-Cell Characterization of Terminal Phenotypes hydroxymate (tRNAser aminoacylation inhibitor). We observed
Provides a Novel View of Essential Gene Function dose-dependent phenotypes (Figure 6C): shorter cells with slight
We examined whether substantial depletion results in more depletion/inhibition (Figures S5C and S5D), filamentation and
extreme morphological changes than basal knockdown by bending with intermediate levels, and growth halting at high
imaging the entire library after prolonged (24 hr) induction of levels; these phenotypes were also exhibited by other amino-
dcas9 (‘‘terminal phenotypes;’’ Supplemental Experimental Pro- acyl-tRNA synthetase knockdowns (Figures 6Aiv, 6Avi, and
cedures). Remarkably, >60% of the strains (174/289 or 166/258 6Aviii). Thus, the rate of inhibition of essential processes can
bona fide essentials) exhibited morphological phenotypes, even qualitatively affect their terminal phenotypes.
excluding strains with only growth defects (overdivision or In summary, the majority of essential genes knockdowns
growth halting; Figure 6Ai), a number far exceeding the 48 known display an altered terminal morphological phenotype. Pheno-
cell envelope-related strains. Knockdowns displayed several types group both within and across functional processes,
predominant morphologies (Figures 6Aii–6Aviii and S5B), demonstrating the utility of morphology for revealing interac-
whereas the control formed a uniform lawn of normal, rod- tions, and can vary with the kinetics and extent of gene
shaped cells (Figure 6Aii). knockdown.

(B) Matrix of the similarity of terminal phenotypes achieved by genes belonging to the most common essential-gene functional groups. b.s., biosynthesis.
(C) Titrated depletion of the valS knockdown and inhibition of WT by serine hydroxamate led to similar trends in terminal phenotypes, from overdivision to fil-
amenting and bending and, finally, growth halting.
(D) Post-induction growth rates of selected strains with different terminal phenotypes. Most strains reduced growth rates by 20%–30% in the first hour, except for
murD and mbl. murD had a larger decrease in growth rate, whereas mbl had a similar growth rate as the no-sgRNA control.
(E) Morphological phenotypes during depletion post-CRISPRi induction. Phenotypes were observable by 40–60 min post-induction.
See also Figures S5 and S6 and Tables S1 and S5.

Cell 165, 1493–1506, June 2, 2016 1501


A B
CRISPRi multiplexing 1

~140X

~1350X
Relative expression
8 sgRNAs per cell
dCas9 0.1

+ 0.01

0.001

ypvA

ysaA
purT

yocJ

yrpD

ysdC
ykwC
pepT
C D
spoVD
pbpA

pbpH
pbpG
pbpC
pbpD
dacC

pbpB

pbpE

pbpX
ponA
dacA
dacB

pbpF
none

dacF

pbpI

1
none
ε (measured - predicted fitness)

dacA 1
dacB 0.8

Measured fitness
dacC
dacF
0.5
pbpA 0.6
pbpB
pbpC
pbpD 0 0.4
pbpE
pbpF pbpD / pbpA / ponA
pbpG 0.2
-0.5
pbpH
pbpI
pbpX 0
ponA -1 0 0.2 0.4 0.6 0.8 1
spoVD N.D. Predicted fitness

E
5 μm

D A A A bpA A A A nA
trol pbp pbp pon pbp / ∆p pon pon pon / po
con D/ D/ A/ A/ bpA
pbp ∆pb
pD pbp pbp / pbp ∆p
p bpD p D/
∆pb
F p < 0.001 G p < 0.000001
Minimum negative curvature

16
(μm-1, abs. value)

1.6
Cell length (μm)

12
1.2

8
0.8

4 0.4

0 0
Cell length
pA
pD onA

A
A
A
A

bp ponA
nA

pA

∆p D / p A
D / bpA
A
A
A
bp ponA
nA
pD

pD
ol

ol
bp / pbp
pb ∆pbp
on
bp / pbp pon

pb ∆pbp
pb / pon
bp / pbp pon
ntr

ntr

Positive curvature
po
pb

pb

po
pb

pb
p

/p

p
co

co

Negative curvature
/
A/
A/

/
A/
A/
pD bpA

pA
pD

pD
D/

p
pb

pb
bp
p

∆p

∆p
∆p

pD
D/

D/
pb

pb
∆p

∆p

(legend on next page)

1502 Cell 165, 1493–1506, June 2, 2016


Cellular Behavior during the Depletion of Essential neously knockdown two genes (Qi et al., 2013); we show that
Proteins CRISPRi can simultaneously, effectively knockdown eight unre-
To determine how rapidly terminal phenotypes are established, lated, nonessential genes (Figures 7A and 7B). Using this multi-
we performed time-lapse imaging on nine representative knock- plexing capability, we examined redundancy for the 16 penicillin
down strains in the presence of inducer. We placed exponen- binding protein (PBP)-encoding genes involved in PG synthesis.
tially growing cells on agarose pads of LB+xylose to initiate Simultaneous knockdown of pairwise pbp genes were viable at
depletion, and imaged them every 5 min to determine microcol- full dcas9 induction, except for pbpA/pbpH, a known synthetic
ony growth rates (Figures 6D and S5E). During the first 60-min lethal pair, and those including pbpB, the only essential PBP
post-induction, WT cells maintained a constant elongation (Figures 7C and S7A–S7C; Table S6) (Wei et al., 2003). The
rate. The mbl knockdown maintained the WT growth rate; observed fitness of double knockdowns was largely predicted
murD had a 50% reduction; and the rest had a 20%–30% by multiplying the fitness of single mutants, a formalism devel-
reduction. Bulging (tagD) was the first prominent morphological oped for null mutations in parallel pathways (Figure S7A; Supple-
phenotype observed (40 min; 1–2 cell doublings on agarose mental Experimental Procedures).
pads) with filamentation (pbpB), bending (accD), and lysis We identified double knockdowns hypersensitive to PBP in-
(murD) occurring after 60 min (2–3 doublings; Figure 6E). As hibitors (mecillinam, cefoxitin, and aztreonam; Figures S7D and
depletion is primarily by dilution, these data indicate that reduc- S7E), as these suggest possible triple synthetic gene combina-
tion of at least 50%–75% below basal knockdown is required to tions. Fitness of the pbpA/ponA/pbpD triple knockdown was
substantially affect growth rate and morphology. Many factors lower than predicted (Figure 7D), and we could not construct a
prevent precise quantification, including differential protein DpbpA/DponA/DpbpD triple deletion, although we could intro-
stabilities and polar effects on transcription. duce an innocuous deletion (DtrpC) or an unrelated pbp deletion
We examined growth of the entire library during dcas9 induc- (DpbpC) into DpbpA/DponA. Thus, the triple deletion is either
tion by diluting stationary-phase cultures into liquid LB+xylose. extremely sick or synthetic lethal.
Approximately one-third of the strains never emerged from sta- PBP4 (pbpD), PBP2a (pbpA), and PBP1a/b (ponA) have trans-
tionary phase (OD600 < 0.06 at 7 hr; Figure S6A). These strains peptidase activity and are septally localized (Scheffers et al.,
covered the entire range of lags, and hence lack of growth could 2004). Reduced septal PG transpeptidation during cell division
not be explained by the variable stationary-phase outgrowth ex- may underlie the severe growth defect of the triple knockdown.
hibited under basal knockdown conditions (Figure S6B). Instead, Control and single- and double-knockdown cells exhibited
these strains identify the subset of essential proteins present in the expected rod-like shape, albeit with reduced cell length (Fig-
limiting amounts such that additional depletion beyond basal ures 7E and 7F), potentially implicating these three genes in the
knockdown results in growth cessation or lysis (Table S4). regulation of cell division. By contrast, the triple mutant had
Among strains that grew, we identified two distinct pheno- distinct phenotypes, including filamentation, cell lysis (Figure 7E,
types (Figure S6C; Table S4; Supplemental Experimental Proce- white arrows), and bending along the cell contour (Figure 7G;
dures). 13 strains showed nearly linear growth; 9/13 affected p < 106), whether constructed with complete knockdown
cofactor biosynthesis (e.g., heme biosynthesis) or electron trans- of all three genes or ponA knockdown in the DpbpA/DpbpD
port (Figure S6D) and may reflect cofactor availability. 14 strains double. This phenotype is consistent with lethality caused by
enriched in either cell-envelope synthesis (5/14), or DNA replica- reduced septal transpeptidation and showcases the ease of dis-
tion (7/14) (Table S4) stalled after initial growth and then exhibited secting redundant pathways with CRISPRi.
a marked decrease in OD, indicative of lysis (Figure S6E). This
common phenotype may reflect the close network and morpho- DISCUSSION
logical connections observed between DNA replication and
envelope synthesis (Figure 3A). Bacteria typically have several hundred essential genes that
encode the core reactions central to viability, together consti-
Dissecting a Highly Redundant Gene Network with tuting 10% of their total genetic complement. Lacking a facile
Multiplexed CRISPRi way to reduce their expression, we had little understanding of
Genetic redundancy, prevalent in complex processes (e.g., con- in vivo relationships among essential gene processes or how
struction of PG; Meeske et al., 2015), can mask the contributions subtle imbalances in essential pathways impacted cellular ho-
of individual genes to essential processes. CRISPRi can simulta- meostasis. This work is a major advance in the study of bacterial

Figure 7. Multiplexed CRISPRi Knockdowns Facilitate Genetic Analysis of Complex Pathways


(A) Schematic of knockdown of eight genes in a single cell. Knockdowns correspond to the genes in (B).
(B) Quantitative PCR of RNA levels in the eight-gene knockdown strain at maximal dcas9 induction (1% xylose). Expression is relative to the no-sgRNA control.
Points are the means of at least three measurements, and error bars are ± 1 SD.
(C) The difference between measured and predicted fitness (ε; Supplemental Experimental Procedures) for each pbp double knockdown based on normalized
colony size at maximal dcas9 induction (1% xylose). Negative and positive ε represent reduced or improved fitness, respectively.
(D) Measured versus predicted fitness of pbp triple knockdowns based on normalized colony size at maximal dcas9 induction (1% xylose).
(E) Terminal phenotypes of pbp knockdown strains. Arrows indicate lysed cells.
(F) Box plots of cell length from pbp knockdowns in (E). n > 100 cells for each strain.
(G) Box plots of cell curvature from pbp knockdowns in (E). n > 100 cells for each strain.
See also Figure S7 and Tables S1 and S6.

Cell 165, 1493–1506, June 2, 2016 1503


essential genes, providing a systematic, unbiased study of their However, the essentiality or redundancy of most cell-wall syn-
phenotypes in vivo using CRISPRi (Figure 1) to obtain facile and thesis proteins has made it difficult to uncover the molecular
precise downregulation. Using chemical genomics profiling mechanisms underlying bacterial cell-shape and size determi-
and high-throughput microscopy, we identified complex pheno- nation. Our CRISPRi knockdown approach allowed us to probe
types, including chemical vulnerabilities (Figure 3), growth and this relationship both under conditions of partial knockdown
shape phenotypes (Figures 4 and 5), and terminal death pheno- (1.5- to 3-fold) and during complete depletion. Partial knock-
types (Figure 6). Together, these revealed a complex web of down identifies outliers exceptionally sensitive to depletion.
connections among essential processes. Given that CRISPR As a consequence, we identified the critical envelope gene ac-
systems are broadly active in bacteria, our approach can be tuators of the response, showed that cell width, as well as
readily extended to other bacterial species, including pathogenic length, is controlled (Figure 5), and determined that cell width
and non-culturable species. montonically varies with the extent of depletion of mur genes
Our essential gene network (Figure 3) reveals numerous inter- (Figures 5D and 5E). The few outliers involved in other pro-
process connections not previously annotated and is highly cesses identify putative regulators of cell-wall synthesis. The
enriched in novel connections among distant processes. Pro- translation and DNA-replication-related outliers may tie the
cess inter-connectivity may provide the cell with mechanisms rate of protein synthesis and DNA replication to cell-wall
for restoring cellular homeostasis in response to transient imbal- growth, possibly using previously unrecognized moonlighting
ances. Some connections make intuitive sense. For example, functions of these proteins. Moreover, since maximal growth
folate is likely linked to replication via its necessity for dTTP syn- rate is not substantially affected across the basal knockdown
thesis. Such correlations readily suggest specific hypotheses strains, our findings indicate that cell size and growth rate
that can be directly tested experimentally; e.g., a particular can be, at least, partially decoupled. Additionally, a graded
DNA polymerase subunit senses dTTP levels. Other connections morphological response to the levels of essential proteins
have no facile explanation and underscore that connections be- may drive physiological heterogeneity, enabling population
tween core processes are abundant and understudied. Our adaptation to dynamic environments.
network likely significantly underestimates connections because In contrast, complete depletion probes the intimate relation-
it is based only on the highest confidence interactions to avoid ship between growth and morphology. Here, cells exhibit a
false positives. Comparable datasets in other organisms will pro- wide array of terminal phenotypes with alterations in both growth
vide the basis for evolutionary studies to investigate the extent to and morphology (Figure 6A). By comparison, only a single
which the logic of these functional circuits is conserved across nonessential E. coli gene deletion (rodZ) has a morphological
organisms. phenotype (Shiomi et al., 2008). Some phenotypes are the result
Our studies suggest that the levels of essential B. subtilis pro- of gradual depletion of function (Figure 6C), which may mimic
teins are higher than necessary to maintain optimal growth, as imbalances that transiently occur due to stochasticity in gene
the vast majority of basal knockdown strains grew indistinguish- expression. Tracing changes at different rates of depletion may
ably from WT in exponential phase (Figure 4B). Thus, either pro- provide clues to the cellular program for morphology. Moreover,
tein levels are set high enough to be robust to small (1.5–3-fold) the dynamics of removing an essential function may be as impor-
decreases in expression, or their levels are maintained by an tant to growth and viability as the presence or absence of such a
elaborate posttranscriptional regulatory system. Moreover, there protein.
is a sufficient excess of essential proteins that most strains (70%) The evolution of essential processes has remained largely
emerge from stationary phase even when their gene products mysterious. Long-term evolution experiments have detected
are depleted during regrowth. The 30% of strains unable to many mutational events in the RNA polymerase complex, cell-
exit stationary phase (Figure S6A) are those whose protein levels wall synthesis, and cell-shape determination (Tenaillon et al.,
are closest to the levels necessary for normal growth. 2012), suggesting a great diversity of molecular adaptations,
In stark contrast, almost all basal knockdowns strains ex- including to essential processes. The apparent excess abun-
hibited increased cell inviability during the exit from the station- dance of most essential proteins (Figures 6D and 6E) suggests
ary phase. Thus, even a small decrease in protein product that new functions could rapidly develop from the existing reper-
increases the vulnerability of cell regrowth (Figure 4A). Impor- toire of proteins without compromising growth, possibly explain-
tantly, there is little overlap between strains ceasing growth ing the prevalence of moonlighting proteins (Huberts and van
soon after depletion and those most vulnerable to inhibition of der Klei, 2010). For example, actin and tubulin homologs have
outgrowth from stationary phase (Figure S6B). This discordance diverse roles in bacteria and eukaryotes (Busiek and Margolin,
suggests that the set of essential genes whose expression level 2015). Phenotypic heterogeneity and variable survival during
is close to that necessary for rapid cell growth in ideal conditions the stationary phase, as observed here (Figure 4), may indicate
is distinct from the set necessary for survival during stationary that stressful environments have shaped the evolutionary history
phase. Speculatively, protein levels may reflect the ecological of essential genes. The complexity of the network (Figure 3A)
niche of B. subtilis in soil, where cells spend much more time suggests context dependence that may be critical for the evolu-
in a non-growing state and must survive long enough to enter tion of essential genes and their interactions; perhaps, certain
the sporulation program with high efficiency. essential genes can be rendered non-essential in environments
Classically, genetic perturbations resulting in morphological such as biofilms. Expansion of our study to varied environments
variation, such as homeotic mutants and variegated maize, may provide a more nuanced view of the instances in which a
provided critical insights into fundamental cellular circuits. particular process is limiting for growth.

1504 Cell 165, 1493–1506, June 2, 2016


EXPERIMENTAL PROCEDURES J.M.P.; DA036858 to J.S.W.; OD017887, R01 DA036858 to L.S.Q.; DP2-
OD006466 to K.C.H.; NIH R01 GM102790 to C.A.G.), HHMI funding (to
CRISPRi Library Design, Cloning, Chemical Screens, and Growth J.S.W.), and NSF funding (MCB-1149328) (to K.C.H.).
Analysis
sgRNAs targeted all putative essential genes (subtiwiki.uni-goettingen.de) and Received: November 10, 2015
those recently identified in our B. subtilis gene knockout library (B.M.K., Revised: February 28, 2016
J.M.P., and C.A.G., unpublished data). sgRNAs were designed to target within Accepted: April 20, 2016
the gene body near the 50 end of the gene on the non-template strand. sgRNA Published: May 26, 2016
libraries were cloned via inverse PCR as previously described (Larson et al.,
2013) and strains were constructed using natural competence transformation. REFERENCES
Chemical screening was performed and chemical-gene scores were calcu-
lated as previously described (Nichols et al., 2011). B. subtilis and S. aureus Arjes, H.A., Kriel, A., Sorto, N.A., Shaw, J.T., Wang, J.D., and Levin, P.A.
UppS proteins were purified using nickel-affinity chromatography, and as- (2014). Failsafe mechanisms couple division and DNA replication in bacteria.
sayed using a Kinetic EnzChek pyrophosphate assay (Life Technologies). Curr. Biol. 24, 2149–2155.
The B. subtilis essential gene network was constructed by calculating all pair- Baba, T., Ara, T., Hasegawa, M., Takai, Y., Okumura, Y., Baba, M., Datsenko,
wise Pearson correlations between sgRNA knockdown strains, randomly K.A., Tomita, M., Wanner, B.L., and Mori, H. (2006). Construction of Escheri-
permuting gene identity relative to chemical-gene scores to generate a back- chia coli K-12 in-frame, single-gene knockout mutants: the Keio collection.
ground distribution, then applying a significance cutoff to the correlations Mol. Syst. Biol. 2, 2006.0008.
based on the 95% confidence interval of the background distribution. Popula-
Blomen, V.A., Májek, P., Jae, L.T., Bigenzahn, J.W., Nieuwenhuis, J., Staring,
tion growth curves were obtained from a microplate reader, and growth infor-
J., Sacco, R., van Diemen, F.R., Olk, N., Stukalov, A., et al. (2015). Gene essen-
mation was extracted by fitting the curves to a Gompertz equation (Zwietering
tiality and synthetic lethality in haploid human cells. Science 350, 1092–1096.
et al., 1990).
Breslow, D.K., Cameron, D.M., Collins, S.R., Schuldiner, M., Stewart-Ornstein,
J., Newman, H.W., Braun, S., Madhani, H.D., Krogan, N.J., and Weissman,
High-Throughput Microscopy
J.S. (2008). A comprehensive strategy enabling high-resolution functional
Images of single cells were acquired after transferring cells from a 96-well plate
analysis of the yeast genome. Nat. Methods 5, 711–718.
onto a large-format agarose pad. Analysis of cellular morphologies was per-
formed using custom MATLAB code. A minimum of 100 cells were analyzed Busiek, K.K., and Margolin, W. (2015). Bacterial actin and tubulin homologs in
for quantitative descriptions of single-cell phenotypes. Terminal phenotypes cell growth and division. Curr. Biol. 25, R243–R254.
were examined by spotting cultures outgrown from stationary phase for two Cao, M., Wang, T., Ye, R., and Helmann, J.D. (2002). Antibiotics that inhibit cell
hours in liquid LB medium, transferred to agarose pads containing LB + 1% wall biosynthesis induce expression of the Bacillus subtilis sigma(W) and sig-
xylose, then imaged after overnight incubation. Time-lapse images were taken ma(M) regulons. Mol. Microbiol. 45, 1267–1276.
in an active-control environmental chamber at 37 C (HaisonTech). Further Christen, B., Abeliuk, E., Collier, J.M., Kalogeraki, V.S., Passarelli, B., Coller,
details of methods are in the Supplemental Experimental Procedures. J.A., Fero, M.J., McAdams, H.H., and Shapiro, L. (2011). The essential genome
of a bacterium. Mol. Syst. Biol. 7, 528.
ACCESSION NUMBERS Czarny, T.L., Perri, A.L., French, S., and Brown, E.D. (2014). Discovery of novel
cell wall-active compounds using PywaC, a sensitive reporter of cell wall
The accession number for the RNA-seq and NET-seq data reported in this stress, in the model gram-positive bacterium Bacillus subtilis. Antimicrob.
paper is GEO: GSE74926. Agents Chemother. 58, 3261–3269.
Defeu Soufo, H.J., Reimold, C., Linne, U., Knust, T., Gescher, J., and Grau-
SUPPLEMENTAL INFORMATION mann, P.L. (2010). Bacterial translation elongation factor EF-Tu interacts and
colocalizes with actin-like MreB protein. Proc. Natl. Acad. Sci. USA 107,
Supplemental Information includes Supplemental Experimental Procedures, 3163–3168.
seven figures, and six tables and can be found with this article online at Elbaz, M., and Ben-Yehuda, S. (2010). The metabolic enzyme ManA reveals a
http://dx.doi.org/10.1016/j.cell.2016.05.003. link between cell wall integrity and chromosome morphology. PLoS Genet. 6,
e1001119.
AUTHOR CONTRIBUTIONS Forsyth, R.A., Haselbeck, R.J., Ohlsen, K.L., Yamamoto, R.T., Xu, H., Trawick,
J.D., Wall, D., Wang, L., Brown-Driver, V., Froelich, J.M., et al. (2002).
Conceptualization, J.M.P., A.C., H.S., L.S.Q., K.C.H., and C.A.G.; Methodol- A genome-wide strategy for the identification of essential genes in Staphylo-
ogy, J.M.P., A.C., H.S., T.L.C., M.H.L., B.M.K., J.S.W., E.D.B., L.S.Q., coccus aureus. Mol. Microbiol. 43, 1387–1400.
K.C.H., and C.A.G.; Investigation, J.M.P., A.C., H.S., T.L.C., M.H.L., S.W., Gilbert, L.A., Horlbeck, M.A., Adamson, B., Villalta, J.E., Chen, Y., Whitehead,
C.H.S.L., E.M., and E.H.W.; Formal Analysis, J.M.P., A.C., H.S., T.L.C., E.H., Guimaraes, C., Panning, B., Ploegh, H.L., Bassik, M.C., et al. (2014).
M.H.L., J.S.H., A.L.S., J.S.W., E.D.B., L.S.Q., K.C.H., and C.A.G.; Writing – Genome-scale CRISPR-mediated control of gene repression and activation.
Original Draft, J.M.P., A.C., H.S., L.S.Q., K.C.H., and C.A.G.; Writing – Review Cell 159, 647–661.
& Editing, J.M.P., A.C., H.S., L.S.Q., K.C.H., and C.A.G.; Funding Acquisition,
Goodman, A.L., McNulty, N.P., Zhao, Y., Leip, D., Mitra, R.D., Lozupone, C.A.,
J.M.P., J.S.W., E.D.B., L.S.Q., K.C.H., and C.A.G.; Supervision, J.M.P.,
Knight, R., and Gordon, J.I. (2009). Identifying genetic determinants needed to
J.S.W., E.D.B., L.S.Q., K.C.H., and C.A.G.
establish a human gut symbiont in its habitat. Cell Host Microbe 6, 279–289.
Gräwert, T., Kaiser, J., Zepeck, F., Laupitz, R., Hecht, S., Amslinger, S., Schra-
ACKNOWLEDGMENTS
mek, N., Schleicher, E., Weber, S., Haslbeck, M., et al. (2004). IspH protein of
Escherichia coli: studies on iron-sulfur cluster implementation and catalysis.
We thank D. Rudner for strains/plasmids, O. Rosenberg for S. aureus DNA, G.
J. Am. Chem. Soc. 126, 12847–12855.
Kritikos and N. Typas for colony-sizing software, and A. Grossman, G.W. Li, N.
Typas, and anonymous reviewers for helpful comments. This work was sup- Hardy, L.W., Finer-Moore, J.S., Montfort, W.R., Jones, M.O., Santi, D.V., and
ported by the UCSF Center for Systems and Synthetic Biology (to L.S.Q.), a Stroud, R.M. (1987). Atomic structure of thymidylate synthase: target for
Stanford Graduate Fellowship (to A.C.), a Stanford Agilent Fellowship (to rational drug design. Science 235, 448–455.
H.S.), a Foundation Grant (from the Canadian Institutes of Health Research) Huberts, D.H.E.W., and van der Klei, I.J. (2010). Moonlighting proteins: an
and a Canada Research Chair to (E.D.B.), the NIH (F32 GM108222 to intriguing mode of multitasking. Biochim. Biophys. Acta 1803, 520–525.

Cell 165, 1493–1506, June 2, 2016 1505


Hunt, A., Rawlins, J.P., Thomaides, H.B., and Errington, J. (2006). Functional Schujman, G.E., Choi, K.H., Altabe, S., Rock, C.O., and de Mendoza, D. (2001).
analysis of 11 putative essential genes in Bacillus subtilis. Microbiology 152, Response of Bacillus subtilis to cerulenin and acquisition of resistance.
2895–2907. J. Bacteriol. 183, 3032–3040.
Larson, M.H., Gilbert, L.A., Wang, X., Lim, W.A., Weissman, J.S., and Qi, L.S. Shiomi, D., Sakai, M., and Niki, H. (2008). Determination of bacterial rod shape
(2013). CRISPR interference (CRISPRi) for sequence-specific control of gene by a novel cytoskeletal membrane protein. EMBO J. 27, 3081–3091.
expression. Nat. Protoc. 8, 2180–2196.
Szklarczyk, D., Franceschini, A., Wyder, S., Forslund, K., Heller, D., Huerta-
Larson, M.H., Mooney, R.A., Peters, J.M., Windgassen, T., Nayak, D., Gross, Cepas, J., Simonovic, M., Roth, A., Santos, A., Tsafou, K.P., et al. (2015).
C.A., Block, S.M., Greenleaf, W.J., Landick, R., and Weissman, J.S. (2014). STRING v10: protein-protein interaction networks, integrated over the tree of
A pause sequence enriched at translation start sites drives transcription life. Nucleic Acids Res. 43, D447–D452.
dynamics in vivo. Science 344, 1042–1047.
Tenaillon, O., Rodrı́guez-Verdugo, A., Gaut, R.L., McDonald, P., Bennett, A.F.,
Meeske, A.J., Sham, L.-T., Kimsey, H., Koo, B.-M., Gross, C.A., Bernhardt,
Long, A.D., and Gaut, B.S. (2012). The molecular diversity of adaptive conver-
T.G., and Rudner, D.Z. (2015). MurJ and a novel lipid II flippase are required
gence. Science 335, 457–461.
for cell wall biogenesis in Bacillus subtilis. Proc. Natl. Acad. Sci. USA 112,
6437–6442. Thomaides, H.B., Davison, E.J., Burston, L., Johnson, H., Brown, D.R., Hunt,
Michna, R.H., Commichau, F.M., Tödter, D., Zschiedrich, C.P., and Stülke, J. A.C., Errington, J., and Czaplewski, L. (2007). Essential bacterial functions
(2014). SubtiWiki-a database for the model organism Bacillus subtilis that links encoded by gene pairs. J. Bacteriol. 189, 591–602.
pathway, interaction and expression information. Nucleic Acids Res. 42, Turner, R.D., Vollmer, W., and Foster, S.J. (2014). Different walls for rods and
D692–D698. balls: the diversity of peptidoglycan. Mol. Microbiol. 91, 862–874.
Moche, M., Schneider, G., Edwards, P., Dehesh, K., and Lindqvist, Y. (1999). van Opijnen, T., Bodi, K.L., and Camilli, A. (2009). Tn-seq: high-throughput
Structure of the complex between the antibiotic cerulenin and its target, beta- parallel sequencing for fitness and genetic interaction studies in microorgan-
ketoacyl-acyl carrier protein synthase. J. Biol. Chem. 274, 6031–6034. isms. Nat. Methods 6, 767–772.
Myoda, T.T., Lowther, S.V., Funanage, V.L., and Young, F.E. (1984). Cloning
Wang, T., Birsoy, K., Hughes, N.W., Krupczak, K.M., Post, Y., Wei, J.J.,
and mapping of the dihydrofolate reductase gene of Bacillus subtilis. Gene
Lander, E.S., and Sabatini, D.M. (2015). Identification and characterization of
29, 135–143.
essential genes in the human genome. Science 350, 1096–1101.
Nichols, R.J., Sen, S., Choo, Y.J., Beltrao, P., Zietek, M., Chaba, R., Lee, S.,
Wei, Y., Havasy, T., McPherson, D.C., and Popham, D.L. (2003). Rod shape
Kazmierczak, K.M., Lee, K.J., Wong, A., et al. (2011). Phenotypic landscape
determination by the Bacillus subtilis class B penicillin-binding proteins
of a bacterial cell. Cell 144, 143–156.
encoded by pbpA and pbpH. J. Bacteriol. 185, 4717–4726.
Overkamp, W., Ercan, O., Herber, M., van Maris, A.J.A., Kleerebezem, M., and
Kuipers, O.P. (2015). Physiological and cell morphology adaptation of Bacillus Winzeler, E.A., Shoemaker, D.D., Astromoff, A., Liang, H., Anderson, K., An-
subtilis at near-zero specific growth rates: a transcriptome analysis. Environ. dre, B., Bangham, R., Benito, R., Boeke, J.D., Bussey, H., et al. (1999). Func-
Microbiol. 17, 346–363. tional characterization of the S. cerevisiae genome by gene deletion and
parallel analysis. Science 285, 901–906.
Peters, J.M., Silvis, M.R., Zhao, D., Hawkins, J.S., Gross, C.A., and Qi, L.S.
(2015). Bacterial CRISPR: accomplishments and prospects. Curr. Opin. Micro- Wolff, M., Seemann, M., Tse Sum Bui, B., Frapart, Y., Tritsch, D., Garcia Estra-
biol. 27, 121–126. bot, A., Rodrı́guez-Concepción, M., Boronat, A., Marquet, A., and Rohmer, M.
Qi, L.S., Larson, M.H., Gilbert, L.A., Doudna, J.A., Weissman, J.S., Arkin, A.P., (2003). Isoprenoid biosynthesis via the methylerythritol phosphate pathway:
and Lim, W.A. (2013). Repurposing CRISPR as an RNA-guided platform for the (E)-4-hydroxy-3-methylbut-2-enyl diphosphate reductase (LytB/IspH)
sequence-specific control of gene expression. Cell 152, 1173–1183. from Escherichia coli is a [4Fe-4S] protein. FEBS Lett. 541, 115–120.

Sanders, G.M., Dallmann, H.G., and McHenry, C.S. (2010). Reconstitution of Xu, L., Sedelnikova, S.E., Baker, P.J., Hunt, A., Errington, J., and Rice, D.W.
the B. subtilis replisome with 13 proteins including two distinct replicases. (2007). Crystal structure of S. aureus YlaN, an essential leucine rich protein
Mol. Cell 37, 273–281. involved in the control of cell shape. Proteins 68, 438–445.
Schaechter, M., Maaloe, O., and Kjeldgaard, N.O. (1958). Dependency on me- Xu, H.H., Trawick, J.D., Haselbeck, R.J., Forsyth, R.A., Yamamoto, R.T.,
dium and temperature of cell size and chemical composition during balanced Archer, R., Patterson, J., Allen, M., Froelich, J.M., Taylor, I., et al. (2010).
grown of Salmonella typhimurium. J. Gen. Microbiol. 19, 592–606. Staphylococcus aureus TargetArray: comprehensive differential essential
Scheffers, D.-J., and Pinho, M.G. (2005). Bacterial cell wall synthesis: new in- gene expression as a mechanistic tool to profile antibacterials. Antimicrob.
sights from localization studies. Microbiol. Mol. Biol. Rev. 69, 585–607. Agents Chemother. 54, 3659–3670.
Scheffers, D.-J., Jones, L.J.F., and Errington, J. (2004). Several distinct local- Zwietering, M.H., Jongenburger, I., Rombouts, F.M., and van ’t Riet, K. (1990).
ization patterns for penicillin-binding proteins in Bacillus subtilis. Mol. Micro- Modeling of the bacterial growth curve. Appl. Environ. Microbiol. 56, 1875–
biol. 51, 749–764. 1881.

1506 Cell 165, 1493–1506, June 2, 2016


Supplemental Figures

Figure S1. Further Characterization of the B. subtilis CRISPRi System, Related to Figure 1
(A) Flow cytometry of cells constitutively expressing rfp, with or without Pxyl-dcas9 and an rfp-targeting sgRNA. RFP levels are relative to the sgRNA- control.
(B) Flow cytometry of cells constitutively expressing rfp, with or without PIPTG-dcas9 and an rfp-targeting sgRNA. RFP levels are relative to the sgRNA- control. rfp
knockdown without IPTG induction is < 15%.
(C) Flow cytometry of cells in which dcas9 was induced by adding the specified concentration of xylose that constitutively express rfp and an rfp-targeting sgRNA
shown as distributions of RFP fluorescence for selected xylose concentrations. RFP levels per cell are relative to the sgRNA- control.
(D) NET-seq of cells maximally induced for dcas9 (1% xylose) and constitutively expressing rfp and an rfp-targeted sgRNA versus cells without an sgRNA. Bars
show the total number of normalized RNA 30 -ends downstream of the sgRNA-dCas9 binding site.
(E) Flow cytometry of cells constitutively expressing an rfp-gfp operon and various sgRNAs targeting gfp at maximal induction of dcas9 (1% xylose). RFP and GFP
levels are relative to the sgRNA- control. The dashed line is y = x. Note that the sgRNAs with low efficacy target the template strand (violet).
(F) Growth curves of Pxyl-dcas9 sgRNA- (red), Pxyl-dcas9 sgRNA+ (green), and B. subtilis 168 cells (blue), with 1% xylose supplemented. Shaded areas are mean ±
std for each strain.

Cell 165, 1493–1506, June 2, 2016 S1


Figure S2. Essential Gene Knockdown Screen Reproducibility and Chemical-Gene Phenotypes, Related to Figures 2 and 3
(A) Reproducibility between replicate colonies for all chemical conditions (R: Pearson’s correlation).
(B) Hierarchical clustering of chemical-gene scores (S-scores; Collins et al., 2006) for all chemical conditions.
(legend continued on next page)

S2 Cell 165, 1493–1506, June 2, 2016


(C) A tetrahydrofolate biosynthesis and utilization cluster defined by knockdown strain sensitivity to the DfrA inhibitor trimethoprim and the Sul inhibitor
sulfamonomethoxine.
(D) A cell division/RNA polymerase cluster defined primarily by sensitivities to cell wall-acting antibiotics (aztreonam, cefoxitin, D-cycloserine, and fosfomycin)
and DNA intercalators (ethidium bromide, and acriflavine).
(E) Concentration-dependent inhibition of purified S. aureus UppS by MAC-0170636. Each point is an individual measurement.
(F) A cluster containing ylaN, isoprenoid biosynthesis, and iron-sulfur cluster biosynthesis genes defined by resistance to cell wall-targeting antibiotics, especially
fosfomycin.

Cell 165, 1493–1506, June 2, 2016 S3


Figure S3. An Essential Gene Network Reveals Numerous Intra- and Inter-process Connections, Related to Figure 3
The B. subtilis essential gene network as depicted in Figure 3A, except showing sub-networks with only one or two connections and gene labels.

S4 Cell 165, 1493–1506, June 2, 2016


A 1
B
DNA condensation/segregation
DNA repair/recombination
True positive rate (TPR)

DNA replication
0.8 RNA polymerase
RNases
TCA cycle
0.6
aminoacyl-tRNA synthetases
biosynthesis of NAD(P)
0.4 biosynthesis of coenzyme A
biosynthesis of fatty acids
biosynthesis of folate
0.2 biosynthesis of heme/siroheme
intra-process biosynthesis of isoprenoids
inter-process biosynthesis of menaquinone
0 biosynthesis of peptidoglycan
0 0.2 0.4 0.6 0.8 1 biosynthesis of phospholipids
False positive rate (FPR) biosynthesis of teichoic acid
biosynthesis/acquisition of aromatic amino acids
C biosynthesis/acquisition of purine nucleotides
biosynthesis/acquisition of pyrimidine nucleotides
cell division
phenotype correlation

divIB ftsZ
0.89

rpoA cell shape


control of sigma factors
electron transport/other
pbpB ftsL rpoB phosphate metabolism
protein secretion
0.57

rpoC rRNA modification and maturation


divIC ftsW ribosomal proteins
ribosome assembly
tRNA modification and maturation
Cell division RNA polymerase transcription factors/other
translation factors STRING
translation/other essential network
sgRNA+ (normalized reads/gene)

D transporters/other
100 0 5 10 15 40 50
# of intra-process connections

10
rpoB/C F
sigW LB LB + 0.5 mM Fe3+
man
1
spoVG
mtlA/D ∆trpC::kan
R = 0.99
0.1
0.1 1 10 100
sgRNA- (normalized reads/gene)
E
∆ylaN::kan
phenotype correlation
0.85

sufC ispF

sufB sufD ipk dxs

ylaN
0.57

sufS ispH
Unknown
Iron-sulfur cluster Isoprenoid
biosynthesis biosnythesis H
G 1
Instantaneous growth rate (h-1)

control
3.5 0.8 essential genes
RFP basal repression

3
0.6

2.5
0.4
2

0.2
1.5

1 0
3.5 5 10 15 18 20 0 0.1 0.2 0.3 0.4 0.5 0.6
Time (h) Optical density (OD600)

(legend on next page)

Cell 165, 1493–1506, June 2, 2016 S5


Figure S4. Further Characterization of the Essential Gene Network and Basal Knockdown Levels, Related to Figures 3 and 4
(A) ROC curves depicting true positive rates and false positive rates for either intra- or inter-process connections between essential operons.
(B) Absolute number of connections between essential operons in the STRING database (blue) or our essential network (red).
(C) Previously uncharacterized connections between cell division and transcription (red lines); these connections are largely due to shared sensitivities of
ftsL/divIC and rpoB/C knockdowns to cell wall-acting antibiotics and DNA intercalators (see also Figure S2D). Black lines are connections between genes in
different operons and gray lines are connections within operons.
(D) RNA-seq of cells with basal expression of dcas9 and an sgRNA targeting rpoB versus cells without an sgRNA. Reads per gene were normalized by RPKM. The
dashed line is y = x. The rpoBC transcript is reduced 40% by the rpoB sgRNA. Outliers and genes of interest are colored as indicated. The operon containing
manP/A-yjdF is abbreviated as ‘‘man.’’
(E) Previously uncharacterized connections among iron-sulfur cluster biosynthesis, isoprenoid biosynthesis, and the gene of unknown function ylaN (red lines).
Black lines are connections between genes in different operons and gray lines are connections within operons.
(F) ylaN is no longer essential in the presence of additional iron. B. subtilis 168 cells were transformed with PCR products that replaced either trpC (B. subtilis 168 is
already mutant for trpC, so the replacement is neutral), or ylaN with a kanamycin-resistance marker.
(G) Basal CRISPRi knockdown of rfp modestly increases during stationary phase.
(H) Culture optical densities (ODs) determine the instantaneous growth rates for all strains. Instantaneous growth rates were calculated from OD measurements in
Figure 4A. Shaded area is mean ± std for sgRNA- control.

S6 Cell 165, 1493–1506, June 2, 2016


Figure S5. Further Characterization of Cellular Dimensions and Terminal Depletion Phenotypes, Related to Figures 5 and 6
(A) Median cell lengths for a subset of strains measured from both phase contrast images and FM4-64 staining. The two measurements are highly correlated. We
note that our strains did not show as strong a propensity for chaining as other B. subtilis strains, and that it was possible to identify the positions of septa from
phase contrast images.
(B) Cells from a subset of strains with representative terminal phenotypes stained with DAPI and FM4-64 to reveal nucleoid structure and the cell membrane,
respectively.
(C) Distribution of cell lengths of the valS knockdown strain in its terminal phenotype in LB versus LB + 0.005% xylose. Xylose-treated cells were significantly
shorter and exhibited the overdivision phenotype.
(D) Distribution of cell lengths of wild-type cells in its terminal phenotype in LB versus LB + 0.6 mg/mL serine hydroxamate. Cells treated with serine hydroxamate
were significantly shorter and exhibited the overdivision phenotype.
(E) Post-induction instantaneous growth rates for selected strains with different terminal phenotypes. Most strains maintained constant growth rates in the first
hour, except for murD, which decreased in growth rate after 40 min.

Cell 165, 1493–1506, June 2, 2016 S7


Figure S6. Essential Gene Depletion Growth Curves, Related to Figure 6
(A–E) Microplate reader growth curves of essential gene knockdown library strains at full induction of dcas9. Set of strains that displayed noteworthy growth
patterns are plotted in each subfigure. (A) The set of strains that never emerged from stationary phase with dcas9 induction. (B) Strains in (A) displayed a range of
apparent lags without dcas9 induction, indicating that cell death during depletion is not correlated to vulnerability in stationary phase. (C) Strains that continued
growing after dcas9 induction. (D) Strains that exhibited slower, almost linear growth curves. (E) Strains that exhibited biphasic behavior, involving initial growth
followed by stalling and a marked decrease in OD.

S8 Cell 165, 1493–1506, June 2, 2016


Figure S7. pbp Knockdown-Specific Examples, Chemical Screens, and Reproducibility, Related to Figure 7
(A) Relationship between double knockdown fitness predicted by the multiplicative model and fitness measured by colony size. The gray, dashed y = x line
represents the ideal relationship between predicted and measured fitness according the multiplicative model; the red line is a linear fit to the data. Strains with two
sgRNAs targeting the same gene were removed for clarity.
(B) Relative fitness defect of the pbpB single knockdown strain based on normalized colony size.
(C) Agreement between normalized colony sizes for reciprocal double knockdown sgRNA pairs (e.g., amyE::sgRNApbpA and thrC::sgRNApbpH versus
amyE::sgRNApbpH and thrC::sgRNApbpA) at maximal induction of dcas9 (1% xylose).
(D) The difference between measured and predicted fitness (ε values) for each pbp double knockdown based on normalized colony size at maximal induction of
dcas9 (1% xylose) and in the presence of the indicated PBP inhibitor.
(E) Agreement between normalized colony sizes for reciprocal double knockdown sgRNA pairs at maximal induction of dcas9 (1% xylose) and in the presence of
the indicated PBP inhibitor.

Cell 165, 1493–1506, June 2, 2016 S9


Cell, Volume 165

Supplemental Information

A Comprehensive, CRISPR-based Functional Analysis


of Essential Genes in Bacteria
Jason M. Peters, Alexandre Colavin, Handuo Shi, Tomasz L. Czarny, Matthew H.
Larson, Spencer Wong, John S. Hawkins, Candy H.S. Lu, Byoung-Mo Koo, Elizabeth
Marta, Anthony L. Shiver, Evan H. Whitehead, Jonathan S. Weissman, Eric D.
Brown, Lei S. Qi, Kerwyn Casey Huang, and Carol A. Gross
EXTENDED EXPERIMENTAL PROCEDURES

Plasmid Construction

Plasmids are available from the Bacillus Genetic Stock Center (http://www.bgsc.org/). dcas9

was PCR-amplified from pdCas9-bacteria (Addgene #44249) using primers containing BamHI-

compatible BsaI sites, digested with BsaI, then ligated into plasmid pAX01 digested with BamHI

to generate pJMP1 (Pxyl-dcas9, ErmR). The rfp-targeting sgRNARR1 was PCR-amplified from

pgRNA-bacteria (Addgene #44251) using primers containing the veg promoter and EcoRI sites,

digested with EcoRI, then ligated into either pDG1662 digested with EcoRI to generate pJMP2

(Pveg-sgRNARR1, CmR), or pDG1731 digested with EcoRI to generate pJMP3 (Pveg-sgRNANT1,

SpcR). Essential gene and pbp library plasmids were generated from either pJMP2 or pJMP3

using inverse PCR as previously described (Hawkins et al., 2015; Larson et al., 2013). uppSBs

was PCR-amplified from B. subtilis 168 genomic DNA using primers containing SalI and NheI

sites, digested with SalI and NheI, then ligated into pDR110 digested with BamHI to generate

pDR-uppSBs (Pspank-uppSBs, SpcR). uppSBs was PCR-amplified from S. aureus N315 genomic

DNA using primers containing HindIII and SphI sites, digested with HindIII and SphI, then

ligated into pDR110 digested with HindIII and SphI to generate pJMP11 (Pspank-uppSSa, SpcR).

Multiple sgRNA plasmids were constructed using BsaI-mediated cloning as previously

described (Hawkins et al., 2015), except for double pbp sgRNA plasmids, which were

constructed by Gibson assembly (Gibson et al., 2009) using the Gibson Assembly master mix

kit (New England Biolabs).

Strain Construction

Strains are available from the Bacillus Genetic Stock Center (http://www.bgsc.org/). All B.

subtilis strains were constructed using natural competence via either a standard or high-

throughput method (Koo et al., in preparation).

Standard Method

  3
3 ml of MC medium (10.7 g/L potassium phosphate dibasic, 5.2 g/L potassium phosphate

monobasic, 20 g/L glucose, 0.88 g/L trisodium citrate dihydrate, 0.022 g/L ferric ammonium

citrate, 1 g/L casein hydrolysate, 2.2 g/L potassium glutamate monohydrate, 20 mM magnesium

sulfate, 150 nM manganese chloride, 20 mg/L tryptophan) were inoculated with a single colony

of B. subtilis and incubated at 37 °C overnight (≥10 h). The overnight culture was diluted to an

OD600 of 0.1 in 10 mL BMK medium (10.7 g/L potassium phosphate dibasic, 5.2 g/L potassium

phosphate monobasic, 20 g/L glucose, 0.88 g/L sodium citrate dihydrate, 0.022 g/L ferric

ammonium citrate, 2.5 g/L potassium aspartate, 10 mM magnesium sulfate, 150 nM manganese

chloride, 40 mg/L tryptophan, 0.05% yeast extract), then grown in a 125 mL flask at 37 °C with

shaking (250 rpm) until cells reached OD600~1.5. 150 µL of culture were then mixed with ≥100

ng of plasmid DNA in a deep 96-well plate, covered with a breathable film, and incubated at

37 °C without shaking for 10 min, then incubated at 37 °C with shaking (900 rpm) for 2 h. After 2

h, cells were plated on LB agar containing selective antibiotics (7.5 µg/mL kanamycin, a

combination of 1 µg/mL erythromycin and 15 µg/mL lincomycin, 6 µg/mL chloramphenicol or

100 µg/mL spectinomycin [by activity]).

High-throughput Method

Double and triple pbp knockdown libraries were constructed using high-throughput

transformation. Individual wells of a deep 96-well plate containing 300 µL of MC medium were

inoculated with single colonies of strains containing Pxyl-dcas9 and an sgRNA targeting one of

the 16 pbp genes; this plate was incubated at 37 °C with shaking (900 rpm) for at least 10 h.

Cells were then diluted to OD600~0.1 in BMK medium, and 25 µL of diluted cells were added to

≥100 ng of plasmid DNA in a shallow, v-bottom 96-well plate, covered with a breathable film,

and incubated at 37 °C in a humidified incubator without shaking for 16 h. 50 µL of LB were

added to each well and the well contents were plated on LB agar containing selective antibiotics.

Flow Cytometry

  4
Strains were grown overnight in LB in deep 96-well plates, and then back-diluted 1:300 into

fresh LB containing the appropriate concentration of xylose to induce expression of dcas9. After

~5 h of growth, cells were diluted 1:300 into phosphate buffered saline, and red fluorescence

levels (B-A laser) were determined using an LSRII flow cytometer (BD Biosciences). Data for at

least 10,000 cells were collected. Median red fluorescence signals were extracted from FCS

files using FlowJo (FlowJo, LLC); error bars are from three biological replicates.

RNA-seq

Cells were grown in LB or LB + 1% xylose to OD600~0.3. RNA extraction and RNA-seq was

performed as previously described (Larson et al., 2014). RNA-seq data was deposited at the

Gene Expression Omnibus (GSE74926).

NET-seq

Cells were grown in LB + 1% xylose to OD600~0.3. NET-seq was performed as previously

described (Larson et al., 2014). NET-seq data was deposited at the Gene Expression Omnibus

(GSE74926).

sgRNA Design

The 20-nucleotide guide sequences for our sgRNA library were designed to be effective and

specific. We first found all GG dinucleotides in the target genome, and extracted the 20 bases

ending one base 5' of that GG as the potential target candidates for the organism. We then

scored the specificity of each 20-nt guide by passing all the guides through Bowtie with the

genome as the reference, then passing each 19-nt suffix through Bowtie, then each 18-nt suffix,

etc. The score for a guide was the shortest suffix for which only a single alignment was found.

Lower specificity score numbers were therefore better. If we never found a unique alignment at

any length, the score was listed as -1, and the guide was avoided. Next, we annotated each

  5
target with any coding regions overlapping that target. We noted the position of the target

relative to the first coding base of the gene. For each gene we chose a target that was as close

as possible to the 5' end of the coding sequence for that gene and that had a low (good)

specificity score.

sgRNA Efficacy Analysis

To test the efficacy of our essential gene knockdown library, we pinned the library onto

rectangular LB agar plates containing 1% xylose to fully induce expression of dcas9. Of the 299

strains in the library, 258 targeted bona fide essential genes as determined by a gold-standard

gene deletion analysis that attempted to replace every open reading frame in the B. subtilis

genome with an antibiotic resistance marker (Koo et al., in preparation). 244 of the 258 (95%)

strains showed a colony size defect of at least 25% (Table S2). We then tested the 14 strains

that failed to show a colony size defect by measuring growth in liquid LB + 1% xylose using a

microplate reader. We found that 5 of the 14 strains showed 50% or less growth compared to

the growth of the no-sgRNA control at mid-log phase (OD600~0.2; Table S2), bringing the

percentage of effective sgRNAs to 97%. Of the nine strains that failed to show growth defects,

five target phage genes; these genes may not be truly essential, as the antibiotic resistance

marker used in the deletion analysis contains a strong promoter that may upregulate adjacent,

toxic phage genes that are normally transcriptionally silent. Alternatively, dCas9 may be acting

to repress phage transcription that is normally toxic (this appears likely for sgRNAs targeting the

SKIN element repressor, sknR/yqaE). In the cases of yezG (a putative antitoxin) and yhdL (anti-

SigM), CRISPRi repression of upstream genes on the same transcript appears to also silence

expression of toxic proteins normally controlled by these gene products. We conclude that

nearly all sgRNAs targeting essential genes showed repression activity.

Quantitative PCR

  6
Overnight LB cultures of B. subtilis were diluted 1:1000 in 2 mL LB + 1% xylose. These liquid

cultures were grown for 6 h at 37 °C. For RNA extraction, 600 µL of culture were mixed with 600

µL of -20 °C methanol and spun down at full speed for 5 min. Once decanted, the RNA was

extracted with the Qiagen RNeasy RNA isolation kit protocol #7. Following RNA isolation,

genomic DNA was eliminated with the Ambion DNA-Free kit per manufacturer’s instructions.

cDNA was synthesized with the Invitrogen SuperScript III First-Strand Synthesis System.

Random hexamers were annealed, and then cDNA was synthesized via SuperScript III RT.

cDNA was diluted 10-fold to be used in conjunction with Agilent Brilliant II SYBR Green QPCR

Master Mix and 0.6 µM final concentration primers. Primers were designed with Primer3 with the

following parameters: 60 °C melting temperature, 200 base pair-amplified region, and amplified

region within the middle of the gene. Control genes employed consisted of sigA, gyrA, and rpoB.

The qPCR included RNA samples that had DNase treatment, as well as RNase treatment, to

verify that no genomic DNA remained. qPCR experiments were conducted in a Stratagene

Mx3005P qPCR System. Three biological replicates were used in each qPCR.

CRISPRi pbp Double and Triple Knockdown Screening

Double and triple knockdown strains containing two sgRNA plasmids that integrate at either

amyE or thrC were constructed using the high-throughput transformation method described

above. Single colony isolates were stored as glycerol stocks in 96-well format. To screen the

library, cells were robotically pinned from glycerol stocks onto rectangular LB agar plates in 384-

colony format using a ROTOR robot (Singer Instruments), then pinned once more to 1536-

colony format. Cells in 1536-colony format were then pinned to LB + 1% xylose plates to fully

induce dcas9 expression. LB plates with PBP inhibitors also contained 1% xylose.. Plates were

imaged using a Powershot G10 camera (Canon) after ~7-9 h of growth at 37 °C, and colony

size was extracted using the “opacity” setting in the Iris software package (Paradis-Bleau et al.,

  7
2014). Colony size was normalized within plates using internal controls that expressed dcas9,

but did not contain sgRNAs.

The fitness (W) of double (or higher-order) deletion strains can be predicted by

multiplying the fitness values of each of the single deletions (deletions of gene x and gene y)

that comprise the double (predicted Wxy = Wx × Wy); this is known as the multiplicative model (St

Onge et al., 2007). Large values of the deviation between the measured fitness and the

predicted fitness of double deletions (ε = measured Wxy – predicted Wxy) indicate a genetic

interaction. Because our PBP genetic interaction screen used CRISPRi knockdowns rather than

gene deletions, it was unclear if the multiplicative rule would apply to our analysis. To test for

agreement between the multiplicative model and our CRISPRi screen data, we plotted predicted

fitness versus measured fitness for all double knockdown strains (Figure S7A). We found a

strong, linear correlation (Pearson’s R = 0.95) between predicted and measured fitness,

suggesting that the multiplicative model can reliably be applied to our CRISPRi knockdown data.

The small deviation we observed from the ideal multiplicative model (ideal fit: y = x; our fit: y =

0.81x+0.13) may be due to the relatively small sample size of our screen (n = 240 double

knockdowns).

MAC-0170636 Screening and UppS Activity Assay

Cloning, Overexpression, and Purification of B. subtilis UppS

The gene uppS (GeneBank sequence NC_000964.3) was cloned into the pET-19b vector

(Novagene) modified to encode an engineered Tobacco Etch Virus (TEV) protease cleavage

site using primers uppS_FWD: CTAGCATATG CTCAACATACTCAAAAATTG and uppS_REV:

CTAGCTACGAG CTAAATTCCGCCAAA. Primers used to generate the construct are listed

below. UppS was over-expressed in E. coli BL21 (Rosetta) using auto-induction (Studier, 2005).

Cells were harvested by centrifugation at 6000 × g for 25 min at 4 °C. Cells were suspended in

lysis buffer containing 50 mM NaH2PO4 (pH 8.0), 300 mM NaCl, 10 mM imidazole, 250 KU

  8
rLysozyme (Novagen) and an EDTA-free protease inhibitor tablet (Roche Diagnostics). Cells

were lysed using a cell disrupter (Constant Systems Limited, Daventry, UK). Lysates were

cleared by centrifugation at 30,000 × g for 30 min at 4 °C. Nickel-affinity chromatography was

performed on the lysates using a 30 mL free-flow gravity column and 5 mL of Ni-NTA agarose

(Sigma). Once loaded, the column was washed with buffer (10 column volumes) containing 50

mM NaH2PO4 (PH 8.0), 300 mM NaCl, 20 mM imidazole. His-tagged UppS was then eluted in

buffer containing 50 mM NaH2PO4 (PH 8.0), 300 mM NaCl, 250 mM imidazole. Elution fractions

were dialyzed overnight against 20 mM Tris-HCl (pH 8.0), 300 mM NaCl, and 1 mM DTT. Once

dialyzed, the elution fraction was concentrated using Amicon Ultra centrifugation filters (10kDa

cut-off) and quantified using a NanoDrop (Thermo Scientific).

Assay for UppS Inhibition by MAC-0170636

A Kinetic EnzCheck pyrophosphate assay (Life Technologies) was used to assess inhibition of

MAC-0170636 in vitro in accordance with the manufacturers guidelines. The IC50 value was

determined in 100 μL reaction volumes using a flat bottom 96-well plate (Costar 3370) in

duplicate. Reactions with varying concentrations of MAC-0170636 were conducted with a final

concentration of 1% DMSO in the presence of 0.2 mM 2-amino-6-mercapto-7-methyl-purine

ribonucleoside (MESG), 0.625 U of purine ribonucleoside phosphorylase (PNP), 0.2 U of

inorganic pyrophosphatase (PyroP), 0.125 μg of purified UppSBs enzyme, 0.82 μM farensyl

pyrophosphate (FPP) (1xKM) and 65 μM isopentenyl pyrophosphate (IPP) (5x KM). For UppSSa,

0.5 μg of purified UppS enzyme, 0.7 μM farensyl pyrophosphate (FPP) (1xKM), and 80 μM

isopentenyl pyrophosphate (IPP) (5x KM) were used in the reaction. Data were fit using GraFit

V5 (Erithacus Software).

CRISPRi Essential Gene Knockdown Screening and Network Construction

Essential Gene Knockdown Screening

  9
sgRNA plasmids were transformed into a strain containing dcas9 (CAG74209) using the

standard method described above. Single colony isolates of each transformation were grown

overnight in LB + 6 µg/mL chloramphenicol, and stored as glycerol stocks at -80 °C in 96-well

plates. Prior to screening, cells were robotically pinned onto rectangular LB agar plates in 384-

colony format, and then pinned once more to 1536-colony format using a ROTOR robot (Singer

Instruments). To screen the essential library, we pinned from rectangular LB agar plates in

1536-colony format to plates containing the indicated concentration of antibiotic or chemical

stress. Each plate contained four technical replicates for each sgRNA strain and each chemical

concentration was replicated at least twice; most concentrations had four replicate plates.

Chemical concentrations were empirically determined by streaking wild-type B. subtilis 168 onto

round agar plates containing chemicals and then visually inspecting the plates for colony size

defects. We used concentrations in our screen that inhibited growth by 50% or less; chemical

concentrations that inhibited growth such that the median colony size after 14 h was less than

the median colony size on LB after 7 h were discarded. Importantly, the effective concentration

of chemical in our screening plates is likely to be lower than the concentration added to the plate

because of chemical breakdown during the two-day period in which plates were dried at room

temperature to reduce problems with colony smearing due to wet plates. Pinned cells were

grown on rectangular plates for 7-14 h at 37 °C, and then imaged using a custom light box with

a Canon Powershot G10 (Canon).

Data Analysis and Network Construction

Colony sizes were extracted from plate images using the Iris software package (Paradis-Bleau

et al., 2014). Spatial effects were normalized using a quadratic function, median and variance of

colony sizes were normalized between plates, and S-scores were computed using previously

developed software (Collins et al., 2006). False discovery rates were computed from S-scores

as previously described (Nichols et al., 2011).

  10
To construct an essential gene network from gene-chemical scores (i.e., S-scores),

these scores were correlated (Pearson correlation) in a pairwise manner for all sgRNA

knockdown strains, resulting in ~90,000 gene-gene correlations. To establish a cutoff for

correlation significance, we estimated the background distribution of correlations by randomly

permuting gene identity relative to the chemical-gene scores 5,000 times, and then used the 95%

confidence interval of this background distribution (after discarding self correlations that equaled

1) to establish a correlation significance threshold of 0.572 (method adapted from Nichols et al.,

2011); this procedure resulted in 412 significant connections between distinct sgRNA

knockdown strains. Network data was visualized using Cytoscape v3.2.1 (Cytoscape

Consortium; Shannon et al., 2003).

Network Comparison to the STRING Database

Essential operons were defined based on manual curation of global expression data (Nicolas et

al., 2012) and functional annotations from the SubtWiki database (http://subtiwiki.uni-

goettingen.de/). For any essential operon with genes in different annotations, the operon was

split into subsets of genes with the same annotation; correlations between genes in the same

operon were eliminated from the analysis. Pairs of essential operons were defined as interacting

if any pair of genes spanning both operons had a correlation above the threshold value 0.572.

The connectivity between two functional groups was defined as the number of essential operon

interactions between the two groups. If two genes within the same essential operon fell into

distinct functional groups, the interaction between them did not contribute to the functional group

connectivity. Similarly, the interaction between operons that have been classified as single

operons in other studies did not contribute to the functional group connectivity.

Comparison to the STRING database was achieved by downloading the Bacillus subtilis

168-specific interactions scores from http://string-db.org/ (Szklarczyk et al., 2015). Scores were

culled to the “experimental” or “database” categories. Genetic interactions in the STRING

  11
database were combined into essential-operon interactions in an equivalent manner to our

chemical genomics data.

Annotation distance analysis between essential gene operons was performed using

annotations from the SubtiWiki online database (http://subtiwiki.uni-goettingen.de/). SubtiWiki

annotations are broken down into four levels: the first level is very general (e.g., metabolism)

and the fourth level is specific (e.g., RNA polymerase). In cases in which there were only three

levels of annotation, the annotation from the third level was also used for the fourth level (Table

S3). Operons with the same annotation at the fourth level (most specific) were given an

annotation distance of 0, while operons with no overlapping annotations were assigned a 4. As

the annotation distance was subject to bias based on the extent of annotation (e.g., whether or

not strains were annotated at the fourth level), we suggest that the distance be interpreted as a

qualitative metric for annotation differences.

High-resolution Liquid Growth Curve Acquisition and Analysis

All strains were grown in deep 96-well plates for 18 h, then back-diluted 1:200 into 200 μL fresh

LB (for induced growth curves, LB was supplemented with 1% xylose, and the overnight

cultures were grown for 26 h) and grown with shaking at 37 °C in an Infinite M200 plate reader

(Tecan) for 7 h. The absorbance at wavelength 600 nm (OD600) was measured at 7.5-min

intervals.

For uninduced growth curves, the natural logarithm of the optical density was fit to the

Gompertz equation (Zwietering et al., 1990) to determine apparent lag time, maximum specific

growth rate, and from that, the doubling time. Each growth curve was fit individually.

For induced growth curves, several criteria were used to classify the curves into different

categories: 1) no growth: OD600<0.06 throughout the measurement; 2) linear growth: final

OD600>0.2, Pearson correlation coefficient >0.98 between time and OD600 after removal of the

lag and saturation regions of the curve, and the slope at the end of the growth phase is >90% of

  12
the slope at the beginning of growth; 3) growth and death: maximum OD600 is >1.2 × the final

OD600, with final OD600>0.06; 4) others: strains that do not fall into any of the previous three

categories.

Viable Counts

Strains were grown in deep 96-well plates for 18 h. The OD of each overnight culture was

measured using a Genesys 20 spectrophotometer (Thermal Scientific). Overnight cultures were

serially diluted in LB, and at each dilution, 100 µL of culture were plated on LB. Plates were

incubated overnight at 37 °C, and the dilutions with ~100 colonies were used to count the

number of colony forming units (CFUs).

Each overnight culture used for CFU counting was also diluted 1:100 in fresh LB, and 1

μL of the diluted culture was spotted onto a pad of 1.5% agarose in fresh LB. The pads were

incubated at 37 °C for 2 h and then imaged. Dead or non-growing cells remained small and

isolated from other cells, and live cells grew into collections of larger cells. The number of

dead/non-growing cells and live cells were manually counted for each strain; a total of at least

200 cells were counted for each strain.

High-throughput Imaging and Analysis

For phase-contrast and fluorescence imaging, cells were imaged on a Nikon Eclipse TE

inverted fluorescence microscope with a 100X (NA 1.40) oil-immersion objective. Images were

collected using an Andor DC152Q sCMOS camera (Andor Technology, South Windsor, CT,

USA) and µManager v1.3 software (Edelstein et al., 2010). For time-lapse imaging, cells were

maintained at 37 °C during imaging with an active-control environmental chamber

(HaisonTech).

Custom MATLAB (MathWorks) image processing code was used to segment cells and

identify active cell contours from phase microscopy (Monds et al., 2014; Ursell et al., 2014). Cell

  13
widths and lengths were calculated using the MicrobeTracker meshing algorithm (Sliusarenko et

al., 2011). For all single-cell quantification, at least 100 cells were analyzed for each strain.

Enrichment of morphological defects in each functional group was identified by conducting a

two-sample KS test between the average length trend line deviations of strains corresponding to

a functional group, and the entire collection background. For pbp strains, cell lengths were

approximated by half of the contour lengths, and cell bending was defined as the minimum

negative curvature of the cell contour.

Terminal Phenotypes

Microscopy

Cells were back-diluted 1:200 from overnight culture into fresh LB and grown at 37 °C for 2 h in

a plate shaker, and then 1 μL of cells was spotted onto a pad of 1.5% agarose in fresh LB

supplemented with 1% xylose. Cells were incubated on the pads overnight at room temperature,

and then imaged. For imaging under antibiotic treatment and xylose titration, the appropriate

concentrations of antibiotics or xylose were added to the agarose pads. For staining, 1 μg/ML

DAPI (Invitrogen) and/or 2 μg/mL FM4-64 (Invitrogen) were added to the pads.

Analysis

Terminal phenotype for each essential gene knockdown was manually classified using an eight-

dimensional phenotype vector encompassing features of lysis, bulging, uniform shape loss,

bending, filamentation, over-division (shorter cells), and level of post-induction growth. Severe

bulging, filamentation, or lysing phenotypes were classified as dominant such that secondary

phenotypes were ignored. A given functional group of essential operons was described by its

fractional composition of terminal phenotypes from essential operons in that group. The cosine

similarity of terminal phenotypes between functional groups was calculated between the

normalized fractional composition vectors.

  14
SUPPLEMENTAL REFERENCES

Collins, S.R., Schuldiner, M., Krogan, N.J., and Weissman, J.S. (2006). A strategy for extracting
and analyzing large-scale quantitative epistatic interaction data. Genome Biol. 7, R63.

Edelstein, A., Amodaj, N., Hoover, K., Vale, R., and Stuurman, N. (2010). Computer control of
microscopes using µManager. Curr. Protoc. Mol. Biol. Ed. Frederick M Ausubel Al Chapter 14,
Unit14.20.

Gibson, D.G., Young, L., Chuang, R.-Y., Venter, J.C., Hutchison, C.A., and Smith, H.O. (2009).
Enzymatic assembly of DNA molecules up to several hundred kilobases. Nat. Methods 6, 343–
345.

Hawkins, J.S., Wong, S., Peters, J.M., Almeida, R., and Qi, L.S. (2015). Targeted
Transcriptional Repression in Bacteria Using CRISPR Interference (CRISPRi). Methods Mol.
Biol. Clifton NJ 1311, 349–362.

Larson, M.H., Gilbert, L.A., Wang, X., Lim, W.A., Weissman, J.S., and Qi, L.S. (2013). CRISPR
interference (CRISPRi) for sequence-specific control of gene expression. Nat. Protoc. 8, 2180–
2196.

Larson, M.H., Mooney, R.A., Peters, J.M., Windgassen, T., Nayak, D., Gross, C.A., Block, S.M.,
Greenleaf, W.J., Landick, R., and Weissman, J.S. (2014). A pause sequence enriched at
translation start sites drives transcription dynamics in vivo. Science 344, 1042–1047.

Monds, R.D., Lee, T.K., Colavin, A., Ursell, T., Quan, S., Cooper, T.F., and Huang, K.C. (2014).
Systematic perturbation of cytoskeletal function reveals a linear scaling relationship between
cell geometry and fitness. Cell Rep. 9, 1528–1537.

Nichols, R.J., Sen, S., Choo, Y.J., Beltrao, P., Zietek, M., Chaba, R., Lee, S., Kazmierczak, K.M.,
Lee, K.J., Wong, A., et al. (2011). Phenotypic landscape of a bacterial cell. Cell 144, 143–156.

Nicolas, P., Mäder, U., Dervyn, E., Rochat, T., Leduc, A., Pigeonneau, N., Bidnenko, E.,
Marchadier, E., Hoebeke, M., Aymerich, S., et al. (2012). Condition-dependent transcriptome
reveals high-level regulatory architecture in Bacillus subtilis. Science 335, 1103–1106.

Paradis-Bleau, C., Kritikos, G., Orlova, K., Typas, A., and Bernhardt, T.G. (2014). A genome-
wide screen for bacterial envelope biogenesis mutants identifies a novel factor involved in cell
wall precursor metabolism. PLoS Genet. 10, e1004056.

Shannon, P., Markiel, A., Ozier, O., Baliga, N.S., Wang, J.T., Ramage, D., Amin, N.,
Schwikowski, B., and Ideker, T. (2003). Cytoscape: a software environment for integrated
models of biomolecular interaction networks. Genome Res. 13, 2498–2504.

Sliusarenko, O., Heinritz, J., Emonet, T., and Jacobs-Wagner, C. (2011). High-throughput,
subpixel precision analysis of bacterial morphogenesis and intracellular spatio-temporal
dynamics. Mol. Microbiol. 80, 612–627.

St Onge, R.P., Mani, R., Oh, J., Proctor, M., Fung, E., Davis, R.W., Nislow, C., Roth, F.P., and
Giaever, G. (2007). Systematic pathway analysis using high-resolution fitness profiling of
combinatorial gene deletions. Nat. Genet. 39, 199–206.

  15
Studier, F.W. (2005). Protein production by auto-induction in high density shaking cultures.
Protein Expr. Purif. 41, 207–234.

Szklarczyk, D., Franceschini, A., Wyder, S., Forslund, K., Heller, D., Huerta-Cepas, J.,
Simonovic, M., Roth, A., Santos, A., Tsafou, K.P., et al. (2015). STRING v10: protein-protein
interaction networks, integrated over the tree of life. Nucleic Acids Res. 43, D447–D452.

Ursell, T.S., Nguyen, J., Monds, R.D., Colavin, A., Billings, G., Ouzounov, N., Gitai, Z., Shaevitz,
J.W., and Huang, K.C. (2014). Rod-like bacterial shape is maintained by feedback between cell
curvature and cytoskeletal localization. Proc. Natl. Acad. Sci. U. S. A. 111, E1025–E1034.

Zwietering, M.H., Jongenburger, I., Rombouts, F.M., and van ’t Riet, K. (1990). Modeling of the
bacterial growth curve. Appl. Environ. Microbiol. 56, 1875–1881.

  16

S-ar putea să vă placă și