<<

www.nature.com/scientificreports

OPEN Comparative transcriptomics of multidrug-resistant baumannii in response to Received: 11 May 2016 Accepted: 5 February 2018 treatments Published: xx xx xxxx Hao Qin1,2,6, Norman Wai-Sing Lo3, Jacky Fong-Chuen Loo1, Xiao Lin1,2, Aldrin Kay-Yuen Yim1,2,7, Stephen Kwok-Wing Tsui4, Terrence Chi-Kong Lau 5, Margaret Ip3 & Ting-Fung Chan 1,2

Multidrug-resistant Acinetobacter baumannii, a major hospital-acquired pathogen, is a serious health threat and poses a great challenge to healthcare providers. Although there have been many genomic studies on the evolution and antibiotic resistance of this species, there have been very limited transcriptome studies on its responses to . We conducted a comparative transcriptomic study on 12 strains with diferent growth rates and antibiotic resistance profles, including 3 fast-growing pan-drug-resistant strains, under separate treatment with 3 antibiotics, namely amikacin, , and . We performed deep sequencing using a strand-specifc RNA-sequencing protocol, and used de novo transcriptome assembly to analyze gene expression in the form of polycistronic transcripts. Our results indicated that genes associated with transposable elements generally showed higher levels of expression under antibiotic-treated conditions, and many of these transposon- associated genes have previously been linked to . Using co-expressed transposon genes as markers, we further identifed and experimentally validated two novel genes of which overexpression conferred signifcant increases in amikacin resistance. To the best of our knowledge, this study represents the frst comparative transcriptomic analysis of multidrug-resistant A. baumannii under diferent antibiotic treatments, and revealed a new relationship between transposons and antibiotic resistance.

Acinetobacter baumannii is a nosocomial pathogen that can grow in diverse environments, including intensive care units. Immunocompromised patients are vulnerable to A. baumannii infection even during therapy. Tese enter the body through moist tissues and can colonize various sites, such as the respiratory tract, cen- tral nervous system, skin, and eyes. Pneumonia, blood stream infections, and are the most frequent adverse efects of infection1. Te major challenge in the treatment of A. baumannii infection is multidrug resist- ance. Pan-drug-resistant A. baumannii strains have been reported across Asia and Europe2–4. Several comparative genomic studies of A. baumannii have discussed the relationships among the variability of drug resistance islands and drug-resistant phenotypes2–4. Multidrug-resistant strains possess a higher metabolic capacity in environments with limited resources5. Nonetheless, information from genome sequences is not suf- cient to capture the dynamics of gene expression. Several transcriptome studies have probed the transcriptional changes in the A. baumannii reference strain ATCC17978 when grown under harsh environmental conditions, such as ethanol treatment6, high salinity7, and iron defciency8. Tese studies particularly noted the strong surviv- ability of A. baumannii under such conditions. Recent transcriptome studies have also investigated the response of this species to antibiotics. For example, a study in which the A. baumannii reference strain ATCC19606 was

1School of Life Sciences, The Chinese University of Hong Kong, Hong Kong SAR, . 2Partner State Key Laboratory of Agrobiotechnology, The Chinese University of Hong Kong, Hong Kong SAR, China. 3Department of Microbiology, The Chinese University of Hong Kong, Hong Kong SAR, China. 4School of Biomedical Sciences, The Chinese University of Hong Kong, Hong Kong SAR, China. 5Department of Biomedical Sciences, City University of Hong Kong, Hong Kong SAR, China. 6Present address: 3D Medicines Corporation, Shanghai, China. 7Present address: Washington University School of Medicine, Saint Louis, MO, USA. Norman Wai-Sing Lo and Jacky Fong-Chuen Loo contributed equally to this work. Correspondence and requests for materials should be addressed to T.-F.C. (email: [email protected])

SCieNTiFiC ReporTs | (2018)8:3515 | DOI:10.1038/s41598-018-21841-9 1 www.nature.com/scientificreports/

treated with and doripenem revealed the critical genes involved in the killing process of colistin9, while in another study the multidrug-resistant A. baumannii strain MDR-ZJ was treated with and identifed an efux pump and several transcriptional regulators involved in the drug resistance response10. Tese studies were performed using only single strains, which may limit the generalizability of their conclusions. Due to the large variations in genome sizes and sequences among diferent strains, especially between multidrug-resistant and -sensitive ones, a more systematic analysis that is free from strain bias is needed to study and compare their transcriptomes under antibiotic-free and -treated environments. Herein, to better understand the transcriptomic response of drug resistance in A. baumannii, we conducted a comparative study of nine multidrug-resistant strains, fve of which are pan-drug-resistant, and three sensitive strains using a strand-specifc RNA-sequencing protocol. We adopted a de novo transcriptome assembly and anal- ysis pipeline to study gene expression in the form of polycistronic transcripts. Diferentially expressed genes were analyzed according to the corresponding strain properties. Tree multidrug-resistant strains were further treated with three antibiotics, namely amikacin, imipenem and meropenem, and the corresponding transcriptomes were generated and analyzed. Our results indicated that genes associated with transposable elements in general were up-regulated under antibiotic treatment. Many of these genes are known to be drug resistance-related. Among several previously unknown transposon-associated genes, we experimentally validated two of them. Expression of these two genes in E. coli BL21 conferred increased resistance to amikacin. In summary, we believe our com- parative transcriptomic study and polycistronic transcript analysis revealed new insights into the drug resistance response in A. baumannii. Our analytical approach should also be applicable to similar studies of other bacterial pathogens. Results A. baumannii strains with diferent drug resistance profles exhibited diferential growth rates in antibiotic-containing media. Twelve strains with diferent pulsed-feld gel electrophoresis (PFGE) pat- terns (Fig. S1) and antibiotic resistance profles were collected (Table S1). Tese include nine multidrug-resistant strains (R1 to R9), fve of which are pan-drug-resistant (R1, R5, R6, R7, and R8), and three sensitive strains (S1 to S3) (Table S1). Seven multidrug-resistant strains (R1 to R7) of the Ia PFGE pattern were clustered to the Global Clone II (GC II)4 of multidrug-resistant A. baumannii (Fig. 1a). Te other strains belonged to clones other than GC I and GC II. Tree multidrug-resistant strains (R7, R8, and R9) grew signifcantly slower than the other nine strains afer 4 hours (Fig. 1b) based on a pooled two-tailed Student’s t-test (p < 0.01).

De novo transcriptome assembly revealed gene-gene relationships in polycistronic transcripts. All 12 strains were cultured in antibiotic-free conditions until mid-log phase. Tree multidrug-resistant strains, R3, R4, and R5, which are from GC II and show similar PFGE patterns, were also cultured in antibiotic-containing media. Tree antibiotics, amikacin (an ), imipenem, and meropenem (both of the class) were chosen, because the 12 strains had exhibited various degrees of resistance to these three antibiotics. In total, 18 antibiotic-treated samples were prepared from either mid-log or stationary phases, but only 17 samples, excluding the amikacin-treated R4 at mid-log phase, were successfully sequenced. RNA sequencing generated a read depth of over 1,000× per sample. A total of 177,939 transcripts were assembled by Trinity11 across 29 samples. 38.8% of the assembled transcripts could be annotated based on known genes from 15 complete A. bau- mannii genomes on the transcribing strand, while the other 61.2% of the transcripts could be classifed into ERCC spike-in sequences, antisense transcripts, or mis-assembled transcripts. Among these annotated transcripts, over 75% were polycistronic, along which more than one gene could be annotated (Fig. S2). Polycistronic mRNAs are produced by prokaryotes to regulate functionally related genes in operons12. In our results, over 97.5% of the genes were annotated from at least one polycistronic transcript (Fig. S3a). Genes that were on the same tran- scripts, particularly those within three loci of one another, had a higher chance of being annotated under the same gene ontology (GO) hierarchy (biological process) compared with those that were on diferent transcripts (p < 0.01) (Fig. S3b). Many of these pairs of genes encoded on the same transcript had signifcantly positive Spearman correlations in terms of expression levels. Te median Spearman correlation between genes on the same transcripts was signifcantly greater than 0 (p < 0.01) (Fig. S3c). Amino acid metabolism and membrane transporters were up-regulated in fast-growing pan-drug- resistant strains. To investigate the major pathways associated with growth rates, the transcriptomes were compared between fast-growing and slow-growing strains at the antibiotic-free mid-log phase. In total, 133 up-regulated and 53 down-regulated genes were identifed in fast-growing strains when compared with slow-growing strains. Based on their GO annotations, six GO terms were functionally correlated with the growth rate based on Pearson’s chi-square test (q < 0.01) followed by the Benjamini–Hochberg procedure (Fig. 1c). Genes that are involved in amino acid metabolic processes and that exhibited the most expression variability among multidrug-resistant strains were all signifcantly up-regulated in fast-growing strains. Contrary to the two slow-growing pan-drug-resistant strains R7 and R8, the two fast-growing strains, R5 and R6, exhibited a high degree of similarity in expression patterns to the fast-growing, sensitive strains S2 and S3, especially for genes involved in amino acid metabolism, glycerol lipid metabolism, siderophore biosynthesis, and transmembrane transporters. Metabolic pathways centered at and around the TCA cycle were consistently up-regulated in all three antibiotic treatments. To investigate which genes and pathways were signifcantly modulated under antibiotic treatment, the transcriptomes of the antibiotic-free mid-log phase were compared with those of the antibiotic-treated mid-log phase and antibiotic-treated stationary phase from three multidrug-resistant strains (R3, R4, and R5). A total of 559 genes were up-regulated and 910 genes were down-regulated in treatment with

SCieNTiFiC ReporTs | (2018)8:3515 | DOI:10.1038/s41598-018-21841-9 2 www.nature.com/scientificreports/

b a R5 1.6 R7 ** ** R6 R4 R1 1.4 R1, R2, R4, S1, S2 R3 R3 R6, S3 R2 Global Clone II R5 BJAB07104 1.2 MDR-TJ MDR-ZJ06 BJAB0868 1 R7, R9 TCDC-AB0715 R8 TYTH-1 1656-2 0.8 ACICU

R8 O.D. 600 R9 0.6 S1 AB0057 AB307-0294 Global Clone I 0.4 AYE S2 n = 3 SDF 0.2 ** P-value < 0.01 ATCC17978 S3 BJAB0715 D1279779 0 123456 0.002 Time (hours)

c LowHigh

G420 G3936 G4166 G1468 G322 G1767 Group III G496 G3350 G1827 G3934 Siderophore biosynthetic G323 G1269 G1081 G3351 G321 process G3707 G3708 G1118 G1192 G980 Group IV G2902 Group I G1533 G4076 G2012 G425 G4239 Transmembrane transporter G4179 Amino acid G1897 G426 G1594 G589 activity G3695 metabolic process G1235 G427 G4169 G4203 G4167 G81 G3928 G4106 G4162 G3802 G3927 G3851 G4164 Group V G3646 G4161 G2144 G3888 G4158 Ribosome biogenesis G1732 G3923 G292 G4171 G362 G4172 G2126 G4184 G2161 Group II G4159 G2266 G1868 G1407 Glycerol lipid G1384 G1625 Group VI G959 metabolic process G1622 G1813 G1619 Proton transporting G4253 G1620 S1 S2 S3 R1 R2 R3 R4 R5 R6 R7 R8 R9 S1 S2 S3 R1 R2 R3 R4 R5 R6 R7 R8 R9 ATP synthase Fast-growing Slow-growing Fast-growing Slow-growing strains strains strains strains

Figure 1. Fast-growing pan-drug-resistant strains exhibited higher nutrition uptake and utilization ability. (a) Phylogenetic tree including collected strains and 15 published complete genomes. (b) Growth curves of 12 strains in BHI broth over 5 hours. Error bars: standard error. p-values were calculated by Student’s t-test. (c) Heat map of diferentially expressed genes between fast-growing strains and slow-growing strains, classifed by functional groups.

at least one of the three antibiotics (Fig. S4a). Among them, 172 genes were commonly up-regulated in all three antibiotic treatments, with genes involved in metabolism constituting the largest group of these (Fig. 2). Genes coding for enzymes involved in the TCA cycle, and in the biosynthesis of some amino acids, purines, and pyri- midines whose raw materials originate from the TCA cycle, were up-regulated. Te complete operons encoding enzyme complexes responsible for ATP synthesis, including NADH dehydrogenase, , and ATP synthase, and ribosomal operons, were also up-regulated. Antibiotic-specifc responses were also studied. Te three A. baumannii strains exhibited distinct responses to the two classes of antibiotics (Fig. S4a). Amikacin specifcally induced up-regulation of 149 genes, but only 4 genes were down-regulated. Tose up-regulated genes were signifcantly enriched in unfolded protein binding, protein folding, and proteolysis based on a hypergeometric test of enrichment (q-value < 0.01), including two protein-folding chaperons (DnaK and HtpX), several peptidyl-prolyl isomerases, and the CLP protease system (Fig. 2). In contrast, more than 400 genes were commonly down-regulated when treated with either of the two carbapenem-related antibiotics, and only 17 genes were commonly up-regulated. Many transcription factors were commonly down-regulated in treatment with imipenem or meropenem. However, up to 60% of the genes that were commonly down-regulated in carbapenem do not have any GO-term annotation. Te three strains also exhibited strain-specifc responses under diferent antibiotic treatments (Fig. S4b). Many more genes were down-regulated in the pan-drug-resistant strain R5 when compared with R3 and R4. About 24% of these R5-specifc down-regulated genes were afected by all three antibiotic treatments, of which many were transmembrane transporters and other membrane involved in signal transductions (Fig. S4c).

SCieNTiFiC ReporTs | (2018)8:3515 | DOI:10.1038/s41598-018-21841-9 3 www.nature.com/scientificreports/

H+

NADH Cytochrome C ATP synthase Dyhydrogenase oxidase

IlvH IlvC (G4490) (G4491) 2-Oxo-butanoate 2-Oxo-butanoate 2-Oxo-butanoate Isoleucine NADH ADP

Unfolded Mature + NAD+ ATP Threonine Glycine Serine Cysteine H Proteins Proteins Pi Peptidyl-prolyl isomerases ThrC PpiC PpiD SecB SurA SlyD (G175) O-phospho- H+ (G888) (G4249) (G2171) (G246) (G2316) Cystathionine Homocysteine Methionine Chaperons L-homoserine DnaK Eno PGK (G1334) (G4516) (G3387) Phosphoenol-pyruvate Glycerate-2P Glycerate-3P Glycerate-1,3P2 down-regulated HtpX Homoserine transcription factors (G4413) CLP protease system IlvC Amikacin ClpA ClpP ClpS ClpX IlvH AdeR family transcription factors: (G65) (G1679) (G64) (G1681) (G4491) Erroneous (G4490) G269 2,3-Dihydroxy-3- Glycerone-P Glyceraldehyde-3P piptides L-Aspartate Pyruvate 2-Acetolactate Methylbutanoate Other proteases 4-Semialdehyde PckG Fda PbpG PrlC (G4376) AraC family transcription factors: Asd (G4561) (G4371) (G3389) (G181) G1887, G3801, G4432 β-D-Fructose-6P Valine Leucine Amino acids 4-Phospho- L-aspartate FadA FadB Alanine (G79) (G78) AsnC family transcription factors: Acetyl-CoA Fatty acids LysC Prephenate G1205, G3869 (G1644) D-Xylulose-5P AraT AcnB AraT (G4468) (G2235) (G4468) L-Aspartate Oxaloacetate Citrate Isocitrate GntR family transcription factors: G298, G504 Ribosome Mqo Mdh PurA ArgG Tyrosine Phenylalanine (G176) (G4034) (G19) (G249) Malate D-Ribulose-5P Lrp family transcription factors:

Adenylo-succinate L-Arginino-succinate 2-oxo-glutarate L-Glutamate L-Glutamine G3613 tRNA mRNA TCA cycle Asd Fumarate (G181) LysR family transcription factors: rRNA PurB G1348, G1558, G2050, G3455, G3782, (G2155) 2-Acetamido-5- CarB SucC SucB oxopentanoate (G2286) D-Ribose-5P G3805, G4152 (G4316) (G4314) S-Succinyl- Succinate Succinyl-CoA Other RNA Ribosomal proteins SucD dihyrolipoamide-E (G4317) MerR family transcription factors: Proline Ornithine ArgG G1315 (G249) Arginine L-Argininosuccinate Citruline RpoA MocR family transcription factors: (G4187) Dihydro- N-Carbamoyl- Carbamoyl- G284 RpoB UDP UMP Orotidine-5P Orotate orotate L-aspartate phosphate (G3905) RpoC NikR family transcription factors: UTP NdK (G3906) CTP (G2128) G2879 RNA ADP ATP PurB PurM PurL PurD PurF (G2155) (G2139) (G2111) (G2426) (G2261) GTP 5-Phospho- IMP FAICAIR AICAR SAICAR CAIR AIR FGAM FGAR GAR PRPP ribosylamine TetR family transcription factors: G1475, G2198, G2251, G2438, G2890, GDP G3668, G3809, G3867

Ribosomal protein operons

RplC RplD RplW RplB RpsS RplV RpsC RplP RpmC RpsQ RplN RplX RplE RpsN RpsH RplF RplR RpsE RpmD RplO SecY RpmJ RpsM RpsK RpsD RpoA RplQ (G2803) (G4157) (G4158) (G4159) (G4160) (G4161) (G4162) (G4164) (G4165) (G4167) (G4168) (G4169) (G4170) (G4171) (G4172) (G4173) (G4175) (G4176) (G4177) (G4180) (G4181) (G4182) (G4183) (G4184) (G4186) (G4187) (G4188)

RpsG FusA TufA tRNA-Trp SecE NusG RplK RplA RplJ RplL RpoB RpoC (G3931) (G3932) (G3919) (G3920) (G3921) (G3922) (G3923) (G3925) (G3927) (G3928) (G3905) (G3906) Operons in metabolisms

RplU RpmA SucB Lpd SucC SucD CLP protease system (G1715) (G1716) (G4314) (G4315) (G4316) (G4317) ClpS ClpA (G64) (G65) PGK Fda (G3388) Cytochrome C oxidase operons (G3387) (G3389) CyoA CyoB CyoC CyoD CyoE RpsF RpsR RplI ClpP ClpX (G1726) (G1727) (G1728) (G1729) (G1730) (G1732) (G1733) (G1734) AlaS LysC (G1679) (G1681) (G356) (G1644)

ATP synthase operons IlvH IlvC AtpI AtpB AtpE AtpF AtpH AtpA AtpG AtpD AtpC (G1617) (G4490) (G4491) (G1615) (G1616) (G1618) (G1619) (G1620) (G1622) (G1623) (G1625) (G1626) Words & arrows in Red: Consistant up-regulated genes in all three antibiotics treatments FadB FadA Words & arrows in Purple: Consistant up-regulated genes only in amikacin treatment NADH dehydrogenase operons (G78) (G79) NuoA NuoB NuoCD NuoE NuoF NuoG NuoH NuoI NuoJ NuoK NuoL NuoM NuoN Words & arrows in Green: Consistant down-regulated genes in two carbapenems treatment (G1660) (G1661) (G1662) (G1663) (G1664) (G1665) (G1666) (G1667) (G1670) (G1672) (G1673) (G1674) (G1675) Words & arrows in Black: Not consistantly di erentially expressed genes

Figure 2. Genes that were diferentially expressed under antibiotic treatment. Red genes or arrows: genes that were consistently up-regulated under treatment with all three antibiotics. Purple genes or arrows: genes that were consistently up-regulated only under treatment with amikacin. Green genes or arrows: genes that were consistently down-regulated only under treatment with the two carbapenem-related antibiotics.

Antibiotic resistance genes were expressed in both antibiotic-resistant and -sensitive strains. Te Comprehensive Antibiotic Resistant Database13 includes a total of 48 known antibiotic resistance-related genes that were identifed in at least one strain (Fig. 3), which include antibiotic inactivation enzymes, multidrug efux pumps, and direct antibiotic targets. Some of the well-documented resistance-related mutations were also identifed in the multidrug-resistant strains, such as the two fuoroquinolone-resistance-conferring mutations: serine-to-leucine at the 83rd residue in the DNA gyrase subunit A, and serine-to-isoleucine at the 80th residue in the DNA topoisomerase IV subunit A14. Moreover, the expression patterns of the drug resistance genes attributed to the phenotype of the antibiotic-resistant strains. For example, beta-lactamase which is an enzyme to break down the structure of beta-lactam and multidrug efux pump which transports the antibiotics out of the cell (Fig. 3) in particular meropenem were highly expressed in R5 and R6 strains (Table S1) and these two strains were resistant to MEM and IPM. Intriguingly, on the other hand, many antibiotic resistance genes were also found to be expressed in the sensitive strains (Fig. 3). On this basis, the presence of antibiotic resistance in the resistant strains may not be solely due to the expression of one particular antibiotic inactivation enzyme or efux pump. Tis phenomenon also suggested the combinational and synergic efects of the drug resistance genes expressed and functioned with each other in antibiotic resistant bacteria.

Transposon-associated antibiotic resistance genes are major contributors to antibiotic resistance. To investigate the driving factors, other than the expression of antibiotic resistance genes, of drug resistance, we also examined the extent to which antibiotic resistance genes appeared in polycistronic transcripts. Known anti- biotic resistance genes were annotated in over 67% of the polycistronic transcripts (Fig. 4a). Over 50% of the other co-transcribed genes could be annotated based on their GO terms. Tese genes are enriched in several functions, including amino acid metabolic processes, cell division, and transposition, revealing the complex transcriptional regulation of the antibiotic response, with signifcant q-values as calculated by the Benjamini–Hochberg proce- dure (q-values < 0.05). To further identify the genes and pathways that contribute the most to antibiotic resistance, diferential gene expression analysis was performed between multidrug-resistant and -sensitive strains at the antibiotic-free mid-log phase. Four groups of genes that were diferentially expressed between multidrug-resistant and -sensitive strains were, as expected, functionally correlated with antibiotic resistance based on Pearson’s chi-square

SCieNTiFiC ReporTs | (2018)8:3515 | DOI:10.1038/s41598-018-21841-9 4 www.nature.com/scientificreports/

Figure 3. Antibiotic resistance genes were widely distributed in resistant and sensitive strains. Each square represents the expression level of a gene (across the row) in a given strain (down the column). Genes are grouped according to protein functions, followed by the type of antibiotic resistance that the gene is associated with.

statistical test (q < 0.01) followed by the Benjamini–Hochberg procedure (Fig. 4b). Responses to antibiotics that involve either many known antibiotic resistance genes or DNA-mediated transposons were both up-regulated in multidrug-resistant strains. Some genes can be assembled with an annotated on the same transcript, and will be referred to as “transposon-associated” throughout this study. To globally study the diference in expression patterns between transposon-associated genes and non-transposon-associated genes, the cumulative density curves of the expressions of both groups of genes were plotted. Gene expressions of three multidrug-resistant strains, namely R3, R4, and R5, with or without antibiotic treatment, were pooled according to the conditions (antibiotic-free mid-log phase, antibiotic-treated mid-log phase, or antibiotic-treated stationary phase) to be analyzed. Te expression distributions of transposon-associated genes did not show signifcant diferences from non-transposon-associated genes at the mid-log phase under antibiotic-free conditions (Fig. 4c). However, transposon-associated genes generally maintained higher expression levels in both mid-log and stationary phases under antibiotic treatment, indicated by the signifcant shif of the curves to the right (higher expression) below the 70th percentile based on Wilcoxon’s rank-sum test (p < 0.01) (Fig. 4d,e). Tis observation also held true for individual genes when the cumulative density curves were separately plotted with diferent antibiotic treatments and strains (Fig. S5). To confrm whether antibiotic resistance had a statistically signifcant correla- tion with transposon-associated genes by the chi-square test, 39 known drug resistance genes were categorized according to their functions, expression patterns, and their co-transcribed genes (Fig. S6a). Transposable ele- ments showed a higher probability of being co-transcribed with drug inactivation enzymes, and their associ- ated drug resistance genes were more likely to be up-regulated in multidrug-resistant strains, indicated by the signifcant p-values (p < 0.01). Tese fndings led to the observation that up-regulated drug resistance genes were preferentially those coding for drug inactivation enzymes (p = 0.007) (Fig. S6b–d). It was also noticeable that all transposon-associated up-regulated drug resistance genes were drug inactivation enzymes (Table S7). Furthermore, comparing the expressions of known drug resistance genes between the antibiotic-free mid-log phase and the antibiotic-treated mid-log phase of each of the three multidrug-resistant strains (R3, R4, and R5) showed that under the antibiotic-treated conditions, transposon-associated resistance genes generally exhibited up-regulation compared with antibiotic-free conditions in all three multidrug-resistant strains, with signifcant p-values (p < 0.01) based on the two-tailed Student’s t-test. Tey also showed greater up-regulation fold changes than non-transposon-associated antibiotic resistance genes, with signifcant p-values in Wilcoxon’s rank-sum test

SCieNTiFiC ReporTs | (2018)8:3515 | DOI:10.1038/s41598-018-21841-9 5 www.nature.com/scientificreports/

a Monocistronic Other genes f Transcripts 57 ** * ** 439 (11.0 %) Biological Processes: (32.7 %) Response to antibiotic; Proteins Response to stress; Polycistronic Proteins without GOs Transcripts with GOs Amino acids metabolic process; 126 2 905 274 Acyl-carrier-protein biosynthetic process; (67.3 %) (24.4 %) (53.0 %) Folic acid biosynthetic process; Cell division;

Transposition, DNA-mediated; ) Other known resistance genes GO enrichment 60 0 (11.6 %) b Low High (q-value < 0.05) G1418 G3795 G2021 G3796 G3797 G1386 G1387 G2992 G3891 G4338 G4241 G2324 G3924 G3868 G4336 G2818 G2790 G3005 G4344 G278 1 G4342 G2792 G2816 G2771 G3902 G3585 G1273 G4285 G2819 G3584 G2910 G3688 G2960 G2958 G2763 G2858 G2859 G2721 G2722 G2508 G2585 S3 -2 S2 Sensitive strains S1

R9 Log(Fold change R8 R7 R6 -4 R5 Resistant strains R4 R3 Transposon Non-transposon Transposon Non-transposon Transposon Non-transposon R2 associated associated associated associated associated associated R1 Group IGroup II LipopolysaccharidePhosphorelay biosynthetic signalproces transduction system Response to antibiotic Transposon, DNA mediated R3 R4 R5 Group III Group IV Amikacin treated *: p-value < 0.05 Imipenem treated **: p-value < 0.01 Meropenem treated

s

c de 100 100 100 p-value = 0.097 p-value < 0.01 p-value < 0.01 90 90 90 80 80 80 70 70 70 60 60 60 50 50 50 40 40 Percentile Percentile 40 Percentile 30 30 30 20 20 20 10 10 10 Transposon-associated 0 0 0 Non-transposon-associated 2 46810 12 14 468101214 46810 12 14 Log Expression Log Expression Log Expression

Figure 4. Transposons were up-regulated in multidrug-resistant strains and their associated genes maintained relatively higher expression in antibiotic-treated environments. (a) Statistics of genes co-transcribed with known antibiotic resistance genes. (b) Heat map of diferentially expressed genes between multidrug-resistant strains and -sensitive strains, classifed by functional groups. (c–e) Comparison of percentile-of-expression values between transposon-associated genes and non-transposon-associated genes at antibiotic-free mid-log phase (c), antibiotic-treated mid-log phase (d), and antibiotic-treated stationary phase (e). Blue line represents transposon-associated genes; red line represents non-transposon-associated genes. p-values were calculated by Wilcoxon’s rank-sum test. (f) Boxplots of fold changes of known antibiotic resistance genes under the antibiotic treatments. Each point represents a log(fold change) value of a known antibiotic resistance gene under a certain antibiotic treatment. p-values were calculated by Wilcoxon’s rank-sum test.

(p < 0.05) (Fig. 4f). However, non-transposon-associated antibiotic resistance genes were not all up-regulated under antibiotic treatment. Teir expression patterns were found to be strain-specifc. Based on the drug resistance profles and transcriptome results of the 12 strains collected in this study, and combined with the 15 published strains with full genome sequences, several candidate resistance genes were identified (Fig. 5). Among these antibiotic-specific resistance genes, beta-lactamase OXA-23 was discov- ered to be transposon-associated, which was previously reported to confer carbapenem resistance in many carbapenem-resistant strains3,4,15–18. In addition, we identifed that two transposon-associated genes, namely, mph, which codes for the macrolide 2′-phosphotransferase homolog, and mel, which codes for the macrolide efux protein homolog, were assembled onto the same transcript downstream of the transposon gene tnpD in all amikacin-resistant strains (R4, R5, R6, R7, and R8), and were also identifed in the genomes of A. baumannii strains MDR-TJ and TYTH-1 (Fig. 6a). Tis combination of genes can be identifed by BLAST on the plasmids in other amikacin-resistant A. baumannii strains, such as BJAB0714 and BJAB0868, and also in other gram-negative bacteria, such as , Providencia stuatii, , , , , and , with nearly full-length coverage at >90% identities. It should be noticed that mph and mel may not be the determinants for amikacin resistance as other antibiotic resistance genes may have an cooperative efect.

Transposon-associated mph and mel conferred amikacin resistance in E. coli. Reverse-transcription PCR (RT-PCR) confrmed that mph and mel were frequently co-transcribed (Fig. 6b), while tnpD and mel were also traceably co-transcribed (Fig. S7a). It is noticed that amplicon 1 could not be amplifed with RT-PCR, which may indicate an assembly error in this contig. Tis is the disadvantage of transcriptome assembly using short reads19. Quantitative PCR further confrmed that mel and mph had similar expression levels (Fig. 6c). To study whether the two candidates conferred amikacin resistance, mel, mph and the positive control aacBI, which is known to confer amikacin resistance20,21, were independently transformed into the amikacin-sensitive E. coli BL21 (Table S3). Teir

SCieNTiFiC ReporTs | (2018)8:3515 | DOI:10.1038/s41598-018-21841-9 6 www.nature.com/scientificreports/

Hypothetical proteins (on the same transcript in R6) n Macrolide 2'-phosphotransferase

G3226G3221G3218G3211G3210G3209G3208G3207G3204G2991G2966 Beta-lactamaseG2828G2826 G2825 MacrolideG2823 OXA-23G2785 effluxMeropene proteinImipenemAmikacim R1 R2 R3 R4 Color Keys R5 R6 Not detectable or not exist R7 R8 Exist (on genomic data) R9 Lowly expressed Candidate genes S1 Expressed S2 S3 Highly expressed Strains in this study

BJAB07104 BJAB0715 BJAB0868 MDR−TJ MDR−ZJ06 Data not available TCDC−AB0715 TYTH−1 Sensitive Antibiotic resistance 1656−2 ATCC17978 Resistant AYE AB0057 AB307−0294 D1279779 ATCC17978 Published genomes SDF

Carbapenem resistant Amikacin resistant Antibiotic related candidates related candidates resistance

Figure 5. Identifcation of novel antibiotic-specifc resistance genes. Each square represents the expression level or the presence of a gene (down the column) in a given strain (across the row). Te resistance of each strain to the three antibiotics is shown on the right panel.

protein expressions were confrmed by western blotting (Fig. S7b). Both mel and mph expression increased the resistance of BL21 under amikacin treatment at diferent concentrations (Fig. 6d). Te MIC of the negative control was 6 μg/ml, while the aacBI, mel and mph transformed strains had MICs around 16 μg/ml. Discussion Comparative genomics studies have been widely conducted to understand the evolution of pathogens toward multidrug resistance. However, knowledge of the global regulation of antibiotic resistance genes is still inade- quate, due to the lack of RNA-seq data and suitable methodologies for comparative transcriptomics. In this study, we collected nine multidrug-resistant and three multidrug-sensitive A. baumannii strains, cultured them with or without antibiotic treatment, and developed a de novo assembly-based method, by which polycistronic tran- scripts, the major form of transcription in bacteria, could be reconstructed, to analyze the comparative transcrip- tomics. Tis novel method of analysis revealed, for the frst time, that the regulation of antibiotic resistance genes was performed at the level of operons. Antibiotic resistance genes were systematically analyzed to determine, frst, with which genes they were co-transcribed, and second, which genes and pathways were signifcantly modulated under treatment with antibiotics. First, our data suggested that under treatment with both classes of antibiotics, genes in several operons that are involved in ATP, RNA, and protein synthesis were simultaneously up-regulated, suggesting that the rates of energy and protein production were generally elevated among multidrug-resistant A. baumannii. Many genes involved in these pathways were also reported to be up-regulated in Wolbachia under doxycycline treatment22. Moreover, the strains also responded diferently to the two classes of antibiotics. Amikacin, an aminoglycoside that targets the translation machinery, induced up-regulation of over 149 genes, among which were genes involved in protein folding and lysis, while only 4 genes were down-regulated. In contrast, the two carbapenem-related antibiotics (imipenem and meropenem) led to a general down-regulation of gene expression, including of a large number of transcription factors. In addition, most of the imipenem- and meropenem-specifc genes encoded either hypo- thetical proteins or proteins without functional annotation. Tus, there remains a large knowledge gap concern- ing our understanding of the transcriptional response under carbapenem treatment. Second, the presence of a single antibiotic resistance gene may not necessarily confer antibiotic resistance in A. baumannii. A wide spectrum of antibiotic resistance genes have been identifed in all strains, including the resistant and sensitive ones (Fig. 3). Some of these genes, e.g. AdeB (G933) (Table S7), a member belonging to the AdeABC efux pump system, have been reported inadequate in development of aminoglycoside resistance23. We have shown in this study that expression of mel and mph could confer amikacin resistance, though the MIC of E. coli strains expressing AacBI, Mel and Mph under amikacin were around 16 μg/ml, which was lower than the MIC of the multidrug-resistant A. baumannii strains (Table S1). Our data suggest that multiple antibiotic resistance genes are necessary to confer antibiotic resistance in clinically isolated multidrug-resistant strains. Tird, our data showed that antibiotic resistance genes that were co-transcribed with transposons gener- ally had higher fold-changes than other resistance genes, which were not associated with transposons, under antibiotic treatment and were more likely to confer antibiotic-specific resistance. Both multidrug-resistant and -sensitive strains possessed a wealth of antibiotic resistance genes. Most antibiotic resistance genes were involved in important operons and were transcribed with other genes. Based on the GO annotations of their co-transcribed genes, antibiotic resistance genes can be classifed into two categories: transposon-associated

SCieNTiFiC ReporTs | (2018)8:3515 | DOI:10.1038/s41598-018-21841-9 7 www.nature.com/scientificreports/

a .5 c RNA-seq qPCR RNA-seq qPCR (RPKM) (Relative expression) (RPKM) (Relative expression) 400 1 1000 0.09 3. 03 350 0.9 900 0.08 0.8 800 300 0.07 .5 0.7 700 0.06 250 0.6 600 0.05 200 0.5 500 400 0.04 2. 02 150 0.4 0.3 300 0.03 100 0.2 200 0.02 Log10 Expression .5 50 0.1 100 0.01 (Reads number per base) 0 0 0 0 R5 R6 R7 R8 S2 S3 R5 R6 R7 R8 S2 S3

1. 01 RNA-seq qPCR RNA-seq qPCR G2819 G2826 Amplicon 1 Transposase TnpD Macrolide protein (Mel) Amplicon 2Amplicon 3 RNA-seq qPCR (RPKM) (Relative expression) G2819 G2826 G2828 450 0.1 Transposase TnpD Macrolide efflux protein Macrolide 2’-phosphotransferase 400 0.09 0.08 (Mel) (Mph) 350 300 0.07 Genomic DNA RNA 0.06 250 b (20 cycles) (30 cycles) 0.05 200 0.04 R5 R6 R7 R8 S2 S3 R5 R6 R7 R8 S2 S3 150 0.03 100 0.02 Amplicon 1 50 0.01 0 0 Amplicon 2 R5 R6 R7 R8 S2 S3 RNA-seq qPCR Amplicon 3 G2828 Macrolide 2’-phosphotransferase d (Mph)

1.6 1.6 4 ug/ml amikacin 1.4 2 ug/ml amikacin 1.4 8 ug/ml amikacin 1.4 1.2 1.2 1.2 1 1 1 Negative control 0.8 Negative control 0.8 Negative control 0.8 AacBI (positive control) AacBI (positive control) O.D 600 AacBI (positive control) O.D 600 0.6 O.D 600 0.6 Mel 0.6 Mel Mel 0.4 Mph 0.4 0.4 Mph Mph 0.2 0.2 0.2

01234567 0 1234567 0 1234567 Time (hours) Time (hours) Time (hours)

Figure 6. Transposon-associated mel and mph were co-transcribed and both contributed to amikacin resistance. (a) Transcript structure of the transposon gene with the two candidates. (b) Reverse-transcription PCR result of connection regions among three genes (mel, mph, and tnpD). (c) Quantitative PCR results for those three genes. (d) Growth of E. coli transformed with the two candidate genes (mel and mph) in diferent amikacin concentrations.

genes, which are co-transcribed with at least one transposon gene, and non-transposon-associated genes. Transposons are among the major players in , which contributes to the evolution of resist- ance islands in A. baumannii genomes24. It has been reported that the insertion of transposable elements may cre- ate active promoters and induce high-level expressions of downstream genes25,26. Moreover, our results revealed that transposable elements could also shape gene expression patterns under antibiotic treatment. Not all trans- posable elements and their associated genes were up-regulated under antibiotic treatment (Fig. S8a,b), but they in general already have higher fold-changes when compared with non-transposon-associated genes. We found that transposon-associated resistance genes had a higher preference to be associated with drug inactivation enzymes, which was also reported recently in a comparative genomic study2. Terefore, transposable elements could serve as an efcient marker for identifying co-transcribed candidate antibiotic resistance genes, through which two novel amikacin resistance genes, namely mel and mph, were reported and experimentally validated in this study. Although the relationships between antibiotic treatment and transcription of resistance genes-associated trans- posons remains unclear, four proteins have been reported as potential regulators involved in transposons expres- sion27,28. In this study, three of these regulators were identifed, including FIS (G2401), H-NS (G3933) and IFH (G4261) (Table S8). It would be of interest to investigate relationships between transposons expression and these regulators as it could serve as an efective way of reversing drug resistance by inhibiting the transcriptional activ- ities of transposons. Materials and Methods Selection of strains. All strains were isolated from patients at hospitals in Hong Kong (Tables S1 and S2). PFGE was performed to identify the strain lineages according to Seifeit et al.29. Te PFGE patterns were clus- tered using the Unweighted Pair Group Method with Arithmetic Mean. Diferent lineages were determined by a similarity cutof of 80%, with 1.5% position tolerance and 1% band optimization. Te minimum inhibitory con- centrations (MICs) of the antibiotics were determined according to Wiegand et al.30. Multilocus sequence typing was performed as described at PubMLST (http://pubmlst.org/abaumannii/). Strains with diferent lineages or antibiotic resistance properties were selected for further study.

Determination of minimum inhibitory concentration (MIC). Te MIC of the strains was determined and interpreted according to the criteria of Clinical and Laboratory Standards Institute (CLSI)31. Briefy, the

SCieNTiFiC ReporTs | (2018)8:3515 | DOI:10.1038/s41598-018-21841-9 8 www.nature.com/scientificreports/

bacteria were grown in 5% blood agar overnight. Bacterial suspension were prepared to a turbidity equivalent to 0.5 McFarland. Bacteria were inoculated to the antibiotic-containing Mueller-Hinton agar plates and incubated at 37 °C for 16 hours and the MICs were read.

Bacteria culture. A. baumannii was frst cultured in BHI broth overnight. 1 ml overnight culture was inoc- ulated to 100 ml fresh BHI broth. Te mid-log (OD600 ~ 1.00) and stationary (OD600 ~ 2.1) samples were har- vested. Te bacteria cultures were incubated at 37 °C with 200 rpm shaking. To determine the optimal antibiotic concentrations for the antibiotic-treated conditions in LB, 2-fold serial dilutions were performed, starting with the MICs as previously determined, to discover the maximum antibiotic concentrations at which bacterial growth was not signifcantly inhibited compared with that in an antibiotic-free medium. Te fnal antibiotic concentrations used, in a range from 1/8 to 1/64 of the MICs, are listed in Table S4.

Library preparation and strand-specifc RNA-sequencing. Ten to 15 ml of bacteria culture was pel- leted by centrifugation at 3,500 g for 10 minutes at room temperature. Two to 4 ml of RNAprotect (Qiagen) was added to the pellet and mixed by vortexing. Te pellet was re-precipitated by centrifugation at 3,500 g for 10 min- utes at room temperature, and the supernatant was discarded. Standard TRIzol (Life Technologies) protocol was then applied to extract total RNAs from the cell pellets. To ensure that the DNA was completely removed, DNase digestion was performed and the total RNA samples were further purifed by acidic phenol–chloroform. mirVana (Life Technologies) was used to deplete tRNAs and other small RNA portions, and the large RNA molecules were retained. Ribosomal RNAs were then removed by Ribo-Zero rRNA removal kits for gram-negative bacteria (Epicentre). External RNA Controls Consortium (ERCC) RNA Spike-in Control Mixes (Ambion) were added to each rRNA-depleted RNA sample according to the user guide. Te sequencing libraries were prepared using a ScriptSeq v2 RNA-seq Library Preparation kit (Epicentre) with rRNA-depleted samples, and all of the libraries were sequenced by Illumina HiSeq 2000 following the strand-specifc sequencing protocol for 100 cycles. All RNA-seq data have been deposited to SRA with the accession SRP075337.

Reads processing and expression calculation. Te frst six bases of reads and the adaptors were frst removed by in-house-developed pipelines. Ten, low-quality sequences were trimmed by Trimmomatic 0.3032 with the following parameters: SLIDINGWINDOW:4:20 and MINLEN:50. Te transcripts were assembled using Trinity r2013-08-1411 with default parameters and specifying–SS_lib_ type as FR. All transcripts were annotated using BLAST based on the 15 full A. baumannii genomes retrieved from the NCBI (Table S5), with both coverage and identity larger than or equal to 0.8. Overlapped annotations on transcripts were further combined if they overlapped each other by at least 70% of their lengths. Based on the gene annotations of the transcripts, the RPKM (reads per kilobase per million mapped reads) values33 were calculated to determine the gene expression levels. Te expression values were normalized by methods previously described34, which can also be found in the supporting documents, under headings Angular based linear regres- sion, Finding the best regression model and Normalizing samples in diferent conditions. Te expression levels of all genes are listed in Table S8.

Diferential expression determination. Diferentially expressed genes between two phenotypical groups of strains (fast-growing vs slow-growing, resistant vs sensitive) were determined by a spatial angular method (refer to the Supplementary methods for a detailed description). Diferentially expressed genes between antibiotic-free conditions and antibiotic-treated conditions were determined by the following set of criteria: (1) to be consid- ered as an up-regulation under antibiotic treatment, the normalized expression value of the gene in the treated sample at mid-log phase must be larger than or equal to 50 RPKM; and to be considered as a down-regulation, the normalized expression value of the gene in the untreated sample at mid-log phase must be larger than or equal to 50 RPKM; (2) for up-regulation under antibiotic treatment, the normalized expression value of the gene in the treated sample at mid-log phase must be at least 2-fold larger than that in the untreated sample; for down-regulation under antibiotic treatment, the normalized expression value of the gene in the untreated sample at mid-log phase must be at least 2-fold larger than that in the treated sample; (3) fnally, for up-regulation under antibiotic treatment, according to (2), the normalized expression value of the gene in the treated sample at sta- tionary phase must be larger than that in the untreated sample at mid-log phase; and for down-regulation, accord- ing to (2), the normalized expression value of the gene in the untreated sample at mid-log phase must be larger than that that in the treated sample at stationary phase. Any gene that was up-regulated in at least two out of three strains and at the same time did not show any down-regulation was considered a common up-regulated gene (and conversely for the common down-regulated genes). However, because the amikacin-treated R4 sample at mid-log phase had failed to be sequenced, only genes up-regulated or down-regulated in both the R3 and R5 strains were used to determine common diferentially expressed genes for amikacin. For Fig. S5A, the criteria used for the amikacin-treated R4 samples were loosened. Genes whose normalized expression values in the untreated sample at mid-log phase were smaller than those in the treated sample at stationary phase were considered to be up-regulated (likewise, down-regulated genes were considered to be those whose normalized expression values in the treated sample at stationary phase were smaller than those in the untreated sample at mid-log phase).

Other bioinformatic analysis. Details of the expression normalization are described in the supporting documents, under headings Angular based linear regression, Finding the best regression model and Normalizing samples in diferent conditions. For phylogenetic analysis, genes that were expressed in the antibiotic-free sam- ples of all strains and could be identifed in all 15 of the complete A. baumannii genomes were selected, whereas transposon-related genes and common drug resistance genes were excluded. Te sequences were aligned sepa- rately using MUSCLE35 3.8.31 and only the aligned regions were then extracted and concatenated. Te phyloge- netic trees were constructed using the neighbor-joining method in MEGA 636. To annotate GO terms to each

SCieNTiFiC ReporTs | (2018)8:3515 | DOI:10.1038/s41598-018-21841-9 9 www.nature.com/scientificreports/

gene, BLASTX searches for all of the genes were performed against the nr protein database with default param- eters, and only the top 50 hits were retained, which were then annotated by Blast2GO37 with default parameters. GO enrichment was measured in terms of q-values, calculated based on a hypergeometric distribution and the Benjamini–Hochberg procedure. Te q-values of the functional correlation of the GOs were determined using a chi-square distribution in a contingency table and with the Benjamini–Hochberg procedure.

Validation of diferentially expressed genes by quantitative PCR (qPCR). Power SYBR Green PCR master mix (Life Technologies) was used according to the standard protocol. A master mix solution was made for each sample. Each 20 μl technical replicate contained 5 ng templates, 375 nM primers (Table S6), and 1× power SYBR Green PCR master Mix. qPCR was performed using an ABI 7500 fast qPCR machine with 40 thermal cycles. Each thermal cycle began with a holding stage at 50 °C for 2 minutes, following by another holding stage at 95 °C for 10 minutes. Ten, the cycling stage started with a denaturing step at 95 °C for 15 seconds, followed by a 1-minute annealing and elongation step at 60 °C. Te top 3 most stably expressed genes at mid-log phase growing in both the antibiotic-free and antibiotic-treated media were used separately as reference genes. ATP-dependent protease (G707) had the smallest variations between biological replications and ftted the trend of RNA-seq. It was therefore chosen as the reference gene in the qPCR validation step.

First-strand cDNA synthesis. Te SuperScript III frst-strand synthesis SuperMix for qRT-PCR (Life Technologies) was used. Te reaction mixture was prepared according to the standard protocol. Te mixture was frst incubated at 25 °C for 10 minutes, followed by incubation at 50 °C for 30 minutes. Te reaction was termi- nated at 85 °C for 5 minutes and cooled over ice. 2 U of E. coli RNase H was added to the mixture and incubated at 37 °C for 20 minutes.

PCR and Sanger sequencing. A 50-μl reaction mixture was prepared for each PCR reaction, including 45 μl Platinum PCR SuperMix (Life Technologies) solution, 5 μM primers (Table S6) and 5 ng templates. Te reaction started with a 2-minute incubation at 94 °C. Twenty cycles were performed with 30 seconds of denatur- ing at 94 °C, 30 seconds of annealing at 57 °C, and 1 minute/kb extension at 72 °C. Te PCR products were purifed by PureLink Quick Gel Extraction and a PCR purifcation Combo Kit (Life Technologies) based on the standard protocol. Te correct sequences of the PCR products were confrmed by Sanger sequencing.

Bacterial transformation. All plasmids were constructed either using a Champion pET101 Directional TOPO Expression Kit following the standard protocol, or by a gene synthesis service from™ GenScript (Nanjing, China)® on pET15b, which is the closest plasmid to pET101 in the sequence (Table S3). 1 to 10 ng plasmids were transformed to BL21 competent cells (Life Technologies) following the standard protocol. Single colonies were selected from the LB agar plate with 50 ug/ml ampicillin. Te DNA of the plasmids was extracted and the correct sequences were confrmed by Sanger sequencing.

Western blot analysis. Bacterial cells were harvested afer 4 hours induction in 1 mM IPTG. Te OD600 values were adjusted to 1.0. All samples were lyzed in bacterial lysis bufer (1% SDS in PBS) and loaded onto 12% SDS-PAGE for western blot analysis. 1:3000 anti-HIS antibodies (GE 27-4710-01) and 1:2000 anti-FLAG antibod- ies (Sigma F3165) in 5% fat-free milk were used as primary antibodies to detect the expression of the candidate resistance genes, engineered with either an HIS-tag (for AacBI and Mel) or a FLAG-tag (for Mph).

Testing of candidate antibiotic resistance genes. Bacteria transformed with the expression vector were incubated at 37 °C in LB at 150 rpm until OD600 was equal to 0.4. 1 mM IPTG was added and the bacte- ria were incubated for 4 hours to induce protein expression. Te culture was diluted to a fnal OD600 of 0.01. Amikacin at the desired concentrations along with 1 mM IPTG were added, and the bacteria were incubated at 37 °C and 150 rpm. Te OD600 values were taken every one hour. References 1. Howard, A. & O’Donoghue, M. Acinetobacter baumannii: an emerging opportunistic pathogen. 3, 243–250 (2012). 2. Adams, M. D. et al. Comparative genome sequence analysis of multidrug-resistant Acinetobacter baumannii. J. Bacteriol. 190, 8053–8064 (2008). 3. Huang, H. et al. Complete genome sequence of Acinetobacter baumannii MDR-TJ and insights into its mechanism of antibiotic resistance. J. Antimicrob. Chemother. 67, 2825–2832 (2012). 4. Zhu, L. et al. Complete genome analysis of three Acinetobacter baumannii clinical isolates in China for insight into the diversifcation of drug resistance elements. PLoS One 8, e66584 (2013). 5. Vallenet, D. et al. Comparative analysis of Acinetobacters: three genomes for three lifestyles. PLoS One 3, e1805 (2008). 6. Camarena, L., Bruno, V., Euskirchen, G., Poggio, S. & Snyder, M. Molecular mechanisms of ethanol-induced pathogenesis revealed by RNA-sequencing. PLoS Pathog. 6, e1000834 (2010). 7. Hood, M. I., Jacobs, A. C., Sayood, K., Dunman, P. M. & Skaar, E. P. Acinetobacter baumannii increases tolerance to antibiotics in response to monovalent cations. Antimicrob Agents Chemother. 54, 1029–1041 (2010). 8. Eijkelkamp, B. A., Hassan, K. A., Paulsen, I. T. & Brown, M. H. Investigation of the human pathogen Acinetobacter baumannii under iron limiting conditions. BMC Genomics 12, 126 (2011). 9. Henry, R. et al. Te transcriptomic response of Acinetobacter baumannii to colistin and doripenem alone and in combination in an in vitro pharmacokinetics/pharmacodynamics model. J. Antimicrob. Chemother. 70, 1303–1313 (2015). 10. Hua, X., Chen, Q., Li, X. & Yu, Y. Global transcriptional response of Acinetobacter baumannii to a subinhibitory concentration of tigecycline. Int. J. Antimicrob. Agents 44, 337–344 (2014). 11. Grabherr, M. G. et al. Full-length transcriptome assembly from RNA-Seq data without a reference genome. Nat. Biotechnol. 29, 644–652 (2011). 12. Kozak, M. Comparison of initiation of protein synthesis in procaryotes, eucaryotes, and organelles. Microbiol. Rev. 47, 1–45 (1983). 13. McArthur, A. G. et al. Te comprehensive antibiotic resistance database. Antimicrob. Agents Chemother. 57, 3348–3357 (2013).

SCieNTiFiC ReporTs | (2018)8:3515 | DOI:10.1038/s41598-018-21841-9 10 www.nature.com/scientificreports/

14. Ruiz, J. Mechanisms of resistance to quinolones: target alterations, decreased accumulation and DNA gyrase protection. J. Antimicrob. Chemother. 51, 1109–1117 (2003). 15. Smith, C. A. et al. Structural basis for carbapenemase activity of the OXA-23 β-lactamase from Acinetobacter baumannii. Chem. Biol. 20, 1107–1115 (2013). 16. Zhou, H. et al. Genomic analysis of the multidrug-resistant Acinetobacter baumannii strain MDR-ZJ06 widely spread in China. Antimicrob. Agents Chemother. 55, 4506–4512 (2011). 17. Zarrilli, R., Pournaras, S., Giannouli, M. & Tsakris, A. Global evolution of multidrug-resistant Acinetobacter baumannii clonal lineages. Int. J. Antimicrob. Agents 41, 11–19 (2013). 18. Chen, C.-C. et al. Genome sequence of a dominant, multidrug-resistant Acinetobacter baumannii strain, TCDC-AB0715. J. Bacteriol. 193, 2361–2362 (2011). 19. Góngora-Castillo, E. & Buell, C. R. Bioinformatics challenges in de novo transcriptome assembly using short read sequences in the absence of a reference genome sequence. Nat. Prod. Rep. 30, 490 (2013). 20. Sarno, R., Ha, H., Weinsetel, N. & Tolmasky, M. E. Inhibition of aminoglycoside 6′-N-acetyltransferase type Ib-mediated amikacin resistance by antisense oligodeoxynucleotides. Antimicrob. Agents Chemother. 47, 3296–3304 (2003). 21. Lin, D. L. et al. Inhibition of aminoglycoside 6′-N-acetyltransferase type Ib by zinc: reversal of amikacin resistance in Acinetobacter baumannii and Escherichia coli by a zinc ionophore. Antimicrob. Agents Chemother. 58, 4238–4241 (2014). 22. Darby, A. C. et al. Integrated transcriptomic and proteomic analysis of the global response of Wolbachia to doxycycline-induced stress. ISME J. 8, 925–937 (2014). 23. Sheikhalizadeh, V. et al. Comprehensive study to investigate the role of various aminoglycoside resistance mechanisms in clinical isolates of Acinetobacter baumannii. J. Infect. Chemother. 23, 74–79 (2017). 24. Mayers, D. L., Lerner, S. A., Ouellette, M. & Sobel, J. D. in Antimicrobial Drug Resistance (ed. Mayers, D. L.) 1, 18 (Humana Press, 2009). 25. Zhang, Z. & Saier, M. H. A novel mechanism of transposon-mediated gene activation. PLoS Genet. 5, e1000689 (2009). 26. Prentki, P., Teter, B., Chandler, M. & Galas, D. J. Functional promoters created by the insertion of transposable element IS1. J. Mol. Biol. 191, 383–393 (1986). 27. Guerin, E. et al. Te SOS Response Controls Integron Recombination. Science 324, 1034–1034 (2009). 28. Cagle, C. A., Shearer, J. E. S. & Summers, A. O. Regulation of the integrase and cassette promoters of the class 1 integron by nucleoid- associated proteins. Microbiology 157, 2841–2853 (2011). 29. Seifert, H. et al. Standardization and interlaboratory reproducibility assessment of pulsed-feld gel electrophoresis-generated fngerprints of Acinetobacter baumannii. J. Clin. Microbiol. 43, 4328–4335 (2005). 30. Wiegand, I., Hilpert, K. & Hancock, R. E. W. Agar and broth dilution methods to determine the minimal inhibitory concentration (MIC) of antimicrobial substances. Nat. Protoc. 3, 163–175 (2008). 31. CLSI. Performance Standards for Antimicrobial Susceptibility Testing; Twenty-Fifh Informational Supplement. CLSI document M100-S25. 35, (2015). 32. Bolger, A. M., Lohse, M. & Usadel, B. Trimmomatic: a fexible trimmer for Illumina sequence data. Bioinformatics 30, 2114–2120 (2014). 33. Mortazavi, A., Williams, B. A., Mccue, K., Schaefer, L. & Wold, B. Mapping and quantifying mammalian transcriptomes by RNA- Seq. Nat. Methods 5, 621–628 (2008). 34. Yim, A. K.-Y. et al. Using RNA-Seq Data to Evaluate Reference Genes Suitable for Gene Expression Studies in Soybean. PLoS One 10, e0136343 (2015). 35. Edgar, R. C. MUSCLE: multiple sequence alignment with high accuracy and high throughput. Nucleic Acids Res. 32, 1792–1797 (2004). 36. Tamura, K., Stecher, G. & Peterson, D. MEGA6: Molecular Evolutionary Genetics Analysis Version 6.0. Mol. Biol. Evol. 30, 2725–2729 (2013). 37. Conesa, A. et al. Blast2GO: a universal tool for annotation, visualization and analysis in functional genomics research. Bioinformatics 21, 3674–3676 (2005). Acknowledgements Tis study was supported by a Health and Medical Research Fund – Infectious Disease (HMRF) 11100512 from the Food and Health Bureau of the Hong Kong Government, a CUHK Direct Grant 4053041 from the university to T.F.C, and in part supported by an Innovation Technology Fund of the Innovation Technology Commission, the Lo Kwee-Seong Biomedical Research Fund and Lee Hysan Foundation. Author Contributions Conceived and designed the experiments: M.I., S.K.W.T., and T.F.C. Performed the experiments: H.Q., N.W.L., and J.F.L. Analyzed the data and interpreted the results: H.Q., X.L., A.K.Y., and T.F.C. Contributed reagents/ materials/analysis tools: M.I., J.F.L., T.C.K.L., and S.K.W.T. Wrote the paper: H.Q. and T.F.C. Additional Information Supplementary information accompanies this paper at https://doi.org/10.1038/s41598-018-21841-9. Competing Interests: Te authors declare no competing interests. Publisher's note: Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional afliations. Open Access This article is licensed under a Creative Commons Attribution 4.0 International License, which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Cre- ative Commons license, and indicate if changes were made. Te images or other third party material in this article are included in the article’s Creative Commons license, unless indicated otherwise in a credit line to the material. If material is not included in the article’s Creative Commons license and your intended use is not per- mitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this license, visit http://creativecommons.org/licenses/by/4.0/.

© Te Author(s) 2018

SCieNTiFiC ReporTs | (2018)8:3515 | DOI:10.1038/s41598-018-21841-9 11