molecular & biochemical parasitologyschnauferlab.bio.ed.ac.uk/pdfs/surve et al mbp 2017.pdf ·...

8
Molecular & Biochemical Parasitology 211 (2017) 57–61 Contents lists available at ScienceDirect Molecular & Biochemical Parasitology Short technical report NADH dehydrogenase of Trypanosoma brucei is important for efficient acetate production in bloodstream forms Sachin V. Surve a,1 , Bryan C. Jensen a , Meredith Heestand a,2 , Muriel Mazet b,3 , Terry K. Smith c , Frédéric Bringaud b,3 , Marilyn Parsons a,d,∗∗,4 , Achim Schnaufer e,,4 a Center for Infectious Disease Research (Formerly Seattle Biomedical Research Institute), 307 Westlake Ave. N., Seattle, WA, 98109, USA b Centre de Résonance Magnétique des Systèmes Biologiques (RMSB), UMR5536, Université de Bordeaux, CNRS, Bordeaux, France c Biomedical Sciences Research Complex, University of St Andrews, North Haugh, St. Andrews KY16 9ST, United Kingdom d Dept. of Global Health, University of Washington, Seattle, WA, 98195, USA e Institute of Immunology & Infection Research and Centre of Immunity, Infection & Evolution, The University of Edinburgh, Edinburgh, EH9 3FL, United Kingdom a r t i c l e i n f o Article history: Received 11 August 2016 Received in revised form 29 September 2016 Accepted 3 October 2016 Available online 4 October 2016 Keywords: Trypanosoma brucei Mitochondrion NDH2 Respiratory complex I NADH:ubiquinone oxidoreductase Acetate a b s t r a c t In the slender bloodstream form, Trypanosoma brucei mitochondria are repressed for many functions. Multiple components of mitochondrial complex I, NADH:ubiquinone oxidoreductase, are expressed in this stage, but electron transfer through complex I is not essential. Here we investigate the role of the parasite’s second NADH:ubiquinone oxidoreductase, NDH2, which is composed of a single subunit that also localizes to the mitochondrion. While inducible knockdown of NDH2 had a modest growth effect in bloodstream forms, NDH2 null mutants, as well as inducible knockdowns in a complex I deficient back- ground, showed a greater reduction in growth. Altering the NAD + /NADH balance would affect numerous processes directly and indirectly, including acetate production. Indeed, loss of NDH2 led to reduced levels of acetate, which is required for several essential pathways in bloodstream form T. brucei and which may have contributed to the observed growth defect. In conclusion our study shows that NDH2 is impor- tant, but not essential, in proliferating bloodstream forms of T. brucei, arguing that the mitochondrial NAD + /NADH balance is important in this stage, even though the mitochondrion itself is not actively engaged in the generation of ATP. © 2016 The Authors. Published by Elsevier B.V. This is an open access article under the CC BY license (http://creativecommons.org/licenses/by/4.0/). Respiratory chains of prokaryotes and eukaryotes are composed of complexes catalyzing oxidation of NADH along with translo- cation of protons across inner mitochondrial (in eukaryotes) or plasma (in prokaryotes) membranes resulting in gradient forma- tion. This gradient in turn drives the ATP synthesis via ATP synthase (complex V). In addition to the five main complexes in mammalian mitochondria involved in energy production [1], some plants, fungi Corresponding author. ∗∗ Corresponding author at: Center for Infectious Disease Research (formerly Seat- tle Biomedical Research Institute), 307 Westlake Ave. N., Seattle, WA, 98109, USA. E-mail addresses: [email protected] (M. Parsons), [email protected] (A. Schnaufer). 1 Present address: University of Pittsburgh, Department of Cell Biology, 3500 Ter- race Street, Pittsburgh, PA, 15261, USA. 2 Present address: Foster School of Business, University of Washington, Seattle, WA, 98195, USA. 3 Present address: Laboratoire de Microbiologie Fondamentale et Pathogénicité (MFP), UMR5234, Université de Bordeaux, CNRS, Bordeaux, France. 4 These authors contributed equally. and parasites have additional enzymes that complement these complexes. Such enzymes include alternative oxidase [2–4] and alternative or Type II NADH dehydrogenases [5]. Trypanosoma brucei long slender bloodstream forms (BF) rely solely on glucose for their energy requirements; the glucose is metabolized primarily via glycolysis [6]. The mitochondria of these cells possess only two of the large complexes associated with the respiratory chain; complex I (NADH:ubiquinone oxidoreductase; cI), and complex V [7–10]. Differentiation into the transmission competent, but non-proliferative short stumpy BF appears to be associated with up-regulation of cI [9–11]. Our previous charac- terization of cI subunits in T. brucei slender BF showed presence of multi-subunit complexes, but as to whether a complete cI is assem- bled remains unclear [12]. Nonetheless, successful deletion of two cI subunits (NUBM and NUKM) proved that electron transfer within cI is not essential in slender BF and that cI does not contribute signif- icantly to NADH dehydrogenase activity in these cells [12]. These findings were surprising because mRNAs of the mitochondrially encoded subunits of cI are preferentially edited in BF to specify http://dx.doi.org/10.1016/j.molbiopara.2016.10.001 0166-6851/© 2016 The Authors. Published by Elsevier B.V. This is an open access article under the CC BY license (http://creativecommons.org/licenses/by/4.0/).

Upload: others

Post on 19-Mar-2020

5 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Molecular & Biochemical Parasitologyschnauferlab.bio.ed.ac.uk/PDFs/Surve et al MBP 2017.pdf · Molecular & Biochemical Parasitology 211 (2017) 57–61 Contents lists available at

S

Na

STa

b

c

d

e

U

a

ARR2AA

KTMNRNA

ocpt(m

t

a

r

W

(

h0

Molecular & Biochemical Parasitology 211 (2017) 57–61

Contents lists available at ScienceDirect

Molecular & Biochemical Parasitology

hort technical report

ADH dehydrogenase of Trypanosoma brucei is important for efficientcetate production in bloodstream forms

achin V. Survea,1, Bryan C. Jensena, Meredith Heestanda,2, Muriel Mazetb,3,erry K. Smithc, Frédéric Bringaudb,3, Marilyn Parsonsa,d,∗∗,4, Achim Schnaufere,∗,4

Center for Infectious Disease Research (Formerly Seattle Biomedical Research Institute), 307 Westlake Ave. N., Seattle, WA, 98109, USACentre de Résonance Magnétique des Systèmes Biologiques (RMSB), UMR5536, Université de Bordeaux, CNRS, Bordeaux, FranceBiomedical Sciences Research Complex, University of St Andrews, North Haugh, St. Andrews KY16 9ST, United KingdomDept. of Global Health, University of Washington, Seattle, WA, 98195, USAInstitute of Immunology & Infection Research and Centre of Immunity, Infection & Evolution, The University of Edinburgh, Edinburgh, EH9 3FL,nited Kingdom

r t i c l e i n f o

rticle history:eceived 11 August 2016eceived in revised form9 September 2016ccepted 3 October 2016vailable online 4 October 2016

eywords:rypanosoma brucei

a b s t r a c t

In the slender bloodstream form, Trypanosoma brucei mitochondria are repressed for many functions.Multiple components of mitochondrial complex I, NADH:ubiquinone oxidoreductase, are expressed inthis stage, but electron transfer through complex I is not essential. Here we investigate the role of theparasite’s second NADH:ubiquinone oxidoreductase, NDH2, which is composed of a single subunit thatalso localizes to the mitochondrion. While inducible knockdown of NDH2 had a modest growth effect inbloodstream forms, NDH2 null mutants, as well as inducible knockdowns in a complex I deficient back-ground, showed a greater reduction in growth. Altering the NAD+/NADH balance would affect numerousprocesses directly and indirectly, including acetate production. Indeed, loss of NDH2 led to reduced levels

itochondrionDH2espiratory complex IADH:ubiquinone oxidoreductasecetate

of acetate, which is required for several essential pathways in bloodstream form T. brucei and which mayhave contributed to the observed growth defect. In conclusion our study shows that NDH2 is impor-tant, but not essential, in proliferating bloodstream forms of T. brucei, arguing that the mitochondrialNAD+/NADH balance is important in this stage, even though the mitochondrion itself is not activelyengaged in the generation of ATP.

© 2016 The Authors. Published by Elsevier B.V. This is an open access article under the CC BY license

Respiratory chains of prokaryotes and eukaryotes are composedf complexes catalyzing oxidation of NADH along with translo-ation of protons across inner mitochondrial (in eukaryotes) orlasma (in prokaryotes) membranes resulting in gradient forma-

ion. This gradient in turn drives the ATP synthesis via ATP synthasecomplex V). In addition to the five main complexes in mammalian

itochondria involved in energy production [1], some plants, fungi

∗ Corresponding author.∗∗ Corresponding author at: Center for Infectious Disease Research (formerly Seat-le Biomedical Research Institute), 307 Westlake Ave. N., Seattle, WA, 98109, USA.

E-mail addresses: [email protected] (M. Parsons),[email protected] (A. Schnaufer).1 Present address: University of Pittsburgh, Department of Cell Biology, 3500 Ter-

ace Street, Pittsburgh, PA, 15261, USA.2 Present address: Foster School of Business, University of Washington, Seattle,A, 98195, USA.3 Present address: Laboratoire de Microbiologie Fondamentale et Pathogénicité

MFP), UMR5234, Université de Bordeaux, CNRS, Bordeaux, France.4 These authors contributed equally.

ttp://dx.doi.org/10.1016/j.molbiopara.2016.10.001166-6851/© 2016 The Authors. Published by Elsevier B.V. This is an open access article u

(http://creativecommons.org/licenses/by/4.0/).

and parasites have additional enzymes that complement thesecomplexes. Such enzymes include alternative oxidase [2–4] andalternative or Type II NADH dehydrogenases [5].

Trypanosoma brucei long slender bloodstream forms (BF) relysolely on glucose for their energy requirements; the glucose ismetabolized primarily via glycolysis [6]. The mitochondria of thesecells possess only two of the large complexes associated with therespiratory chain; complex I (NADH:ubiquinone oxidoreductase;cI), and complex V [7–10]. Differentiation into the transmissioncompetent, but non-proliferative short stumpy BF appears to beassociated with up-regulation of cI [9–11]. Our previous charac-terization of cI subunits in T. brucei slender BF showed presence ofmulti-subunit complexes, but as to whether a complete cI is assem-bled remains unclear [12]. Nonetheless, successful deletion of twocI subunits (NUBM and NUKM) proved that electron transfer withincI is not essential in slender BF and that cI does not contribute signif-

icantly to NADH dehydrogenase activity in these cells [12]. Thesefindings were surprising because mRNAs of the mitochondriallyencoded subunits of cI are preferentially edited in BF to specify

nder the CC BY license (http://creativecommons.org/licenses/by/4.0/).

Page 2: Molecular & Biochemical Parasitologyschnauferlab.bio.ed.ac.uk/PDFs/Surve et al MBP 2017.pdf · Molecular & Biochemical Parasitology 211 (2017) 57–61 Contents lists available at

58 S.V. Surve et al. / Molecular & Biochemical Parasitology 211 (2017) 57–61

Fig. 1. NDH2 is advantageous but not essential for growth of BF T. brucei in vitro. (A) Southern blot analysis of the NDH2 locus. Genomic DNA was isolated from the parental‘wild type’ (WT) line and ndh2 cKO clone 10. Following digestion with SnaBI and probing with the NDH2 CDS, the expected band sizes are: WT, 3.6 kbp; cKO, >5.5, 2.4,and 2.3 kbp. As the largest fragment in the cKO cell line represents integration of the plasmid bearing the complementing gene into one of the rDNA loci, it is not possibleto accurately predict the fragment size. The lanes are from the same gel and hybridization. (B) Cumulative growth curves of tetracycline (Tet)-treated (induced, Ind) anduninduced (UI) ndh2 cKO cells is compared to the parental WT. For uninduced (UI) cells Tet was removed from the medium at time point zero. Two clones (5 and 10) wereanalyzed. This experiment was performed in duplicate and 91% of the replicates on days 1–4 were within 15% of the mean. (C) Assessment by immunoblot of levels of ectopic,V5-tagged NDH2 in uninduced (-Tet) vs. induced cells (+Tet) on day 5. Gels (5 × 106 cell equivalents per lane) were transferred to nitrocellulose membranes, blocked andincubated with mouse anti-V5 monoclonal antibody at 0.5 �g/ml. Anti-V5 was detected by goat anti-mouse IgG-IRDye 800CW using a Li-Cor Odyssey system. NDH2-V5 has apredicted molecular weight of ∼60 kDa. Lanes are from the same scan of the western blot. (D) Southern analysis confirming the ndh2 cKO genotype in the �nubm background.Genomic DNA digested with SnaBI was analyzed by Southern blot using a probe consisting of the NDH2 coding sequence. Expected sizes are: WT, 3.6 kbp; �nubm ndh2 cKOclones 4 and 5, >5.6 kbp (the complementing copy is in one of the rDNA loci, so its exact size cannot be predicted). The NUBM knockout was confirmed in a previous study[12]. (E) Cumulative growth in the presence and absence of the inducer Tet of parasites with all endogenous alleles of NUBM and NDH2 deleted and harboring an inducible,V5-tagged copy of NDH2. For uninduced (UI) cells Tet was removed from the medium at time point zero. Two individual clones, 4 and 5, were analyzed. Error bars mark thestandard deviation of the triplicate data points. For both clones, the calculated doubling times were ∼9.3 h in the presence of Tet and >100 h in the absence of Tet (days 1–3).( paral( movao

fkadldttTNw

aaa

F) Western analysis of NDH2-V5 expression in the cKOs upon Tet withdrawal, inPGK) as a loading control. The return to normal growth rates 3–4 days after Tet rether experiments (see Fig. S2).

unctional proteins. As at least two NAD+-dependent activities arenown to be essential in BF (the glycine cleavage complex [13] andcetate production via pyruvate dehydrogenase or threonine dehy-rogenase [14]), we reasoned that other enzymes in cI deficient

ines either replace or complement cI’s NADH:ubiquinone oxidore-uctase activity. The type II NADH dehydrogenase NDH2 appearedo be the most likely candidate, as the enzyme can transfer elec-rons from NADH to ubiquinone and was reported to be active in. brucei − at least in the procyclic insect form (PF) [15–17]. Thus,DH2 would be capable of regenerating sufficient NAD+ for useithin the mitochondrion.

T. brucei NDH2 belongs to class A NDH2 enzymes, whichre present in all three domains of life. The T. brucei enzyme,

single polypeptide of ∼54 kDa [16], utilizes a non-covalentlyttached FMN as a cofactor and was proposed to be the source

lel with panel E. The same blot was re-probed with anti-phosphoglycerate kinasel is most likely due to loss of repression of the ectopic NDH2 gene, as was seen in

of rotenone-insensitive NADH dehydrogenase activity in sucrosegradient fractions of PF lysates [15]. Using RNAi to target NDH2(Tb927.10.9440) in PF yielded slower growth and decreasedmitochondrial membrane potential [17]. However, NADH:Q2 oxi-doreductase activity did not change significantly in these NDH2knockdown cells [17]. The authors also proposed that the enzymewas facing the mitochondrial intermembrane space and not thematrix, contrary to the earlier publication [15]. Although presenceof the NDH2 protein in slender BF was confirmed in recent pro-teomic studies [18,19], its physiological role in BF is not known.

To further understand the role of NDH2 in slender BF, we gen-

erated and analyzed the effect of NDH2 knockout (or conditionalknockout, cKO) in wild type and cI deficient lines using T. b. bruceiBF strain Lister 427. We first tested for essentiality by attempt-ing to generate NDH2 null parasites in the single-marker derivative
Page 3: Molecular & Biochemical Parasitologyschnauferlab.bio.ed.ac.uk/PDFs/Surve et al MBP 2017.pdf · Molecular & Biochemical Parasitology 211 (2017) 57–61 Contents lists available at

S.V. Surve et al. / Molecular & Biochemical Parasitology 211 (2017) 57–61 59

Fig. 2. Metabolic changes after NDH2 ablation. (A) Ablation of NDH2 decreases the amount of acetate excreted by the parasites. The amount of acetate, pyruvate and alanineexcreted by the WT cells and the �nubm and/or ndh2 mutant cell lines (given as nmole/h/mg of protein) was determined by 1H NMR spectrometry as previously described[14]. Three biological replicates were performed for each cell line, except for WT (7 replicates) and �nubm (8 replicates). The standard deviations for each condition areindicated. �ndh2 clone a3-10 and �nubm ndh2 cKO clone 4 were used. (B) Characterisation and relative quantification of the total fatty acids present in the lipid extracts ofduplicate biological replicates of each of the various clones. This was done by base hydrolysis and subsequent conversion of the fatty acids to their methyl esters allowinganalysis by GC–MS as previously described [23].

Page 4: Molecular & Biochemical Parasitologyschnauferlab.bio.ed.ac.uk/PDFs/Surve et al MBP 2017.pdf · Molecular & Biochemical Parasitology 211 (2017) 57–61 Contents lists available at

6 hemic

odaMiSdttdsAs

pecVaa(pt(f

swiltcdTbncctlpfeFpotclii

ttcdecpotrpwab

0 S.V. Surve et al. / Molecular & Bioc

f strain Lister 427 [20], using deletion constructs where therug resistance genes were flanked by regions directly upstreamnd downstream of the NDH2 coding sequence (see Supplementalethods for details and primers). Deletion of NDH2 in the result-

ng transfectants was confirmed by PCR (not shown) and genomicouthern analysis (Fig. S1A). In general, we observed a clear growthefect early after transfection, but parasites reproducibly were ableo partially compensate to differing extents upon continued cul-ure. For example, at 46 days in culture, the two knockout linesiffered in their growth characteristics, with one showing slightlylowed growth and the other a much stronger decrease (Fig. S1B).necdotally, we noticed that both of these clones appeared to beensitive to stress, such as recovery from frozen stocks.

Given the clone-specific differences in the �ndh2 parasites, andartial recovery of growth rates over time, we generated cKOs. Thendogenous NDH2 genes were deleted in parasites bearing a tetra-ycline (Tet)-regulated ectopic copy of NDH2 (tagged with three5 epitopes) (see Southern analysis, Fig. 1A). Removal of Tet wasccompanied by slowed growth (the doubling time increased bypproximately 1.5 fold), but the parasites continued to proliferateFig. 1B). Western analysis confirmed the knockdown of the ectopicrotein (Fig. 1C). The C-terminal V5 tags did not interfere with func-ion since induced cells showed growth rates similar to wild typeWT) cells. Thus, NDH2 appears to be beneficial, but not essentialor in vitro growth of slender BF T. brucei.

In earlier work, we showed that slender BF parasites lacking cIubunits NUBM or NUKM,which are required for electron transferithin cI, have no growth defect in vitro or in vivo [12]. However, it

s possible that normal levels of NDH2 are sufficient to fulfill cellu-ar requirements for regeneration of NAD+ in the absence of cI. Weherefore generated ndh2 mutants in the previously characterizedI-deficient line, �nubm [12]. In several attempts, we obtained ndh2ouble knockouts only in the presence of an ectopic copy of NDH2.he ndh2 cKOs in the �nubm parasites were confirmed by Southernlot analysis (Fig. 1D) and by PCR (not shown). A strong growth phe-otype was observed upon NDH2 knockdown in parasites lackingI function; after withdrawal of Tet, their growth slowed dramati-ally, nearly ceasing within 24 h (Fig. 1E). Growth began to recoverhree days to four days after Tet withdrawal, most likely due tooss of repression, as is common in T. brucei (Fig. S2). The growthhenotype of �nubm ndh2 cKO cells was stronger than that seenor ndh2 cKO clones (although NDH2 repression appeared to beven more stringent in the latter (compare Fig. 1C with Fig. 1F andig. S2), and additionally appeared to be stronger than the growthhenotype of the �ndh2 clones. One possible explanation for ourbservation is that cI can partially compensate for NDH2 loss, whichhen would suggest that the two activities function in the sameompartment, the mitochondrial matrix, consistent with an ear-ier report [15]. However, another study has suggested that NDH2s localized to the intermembrane space of the mitochondrion innsect stage parasites [17].

To further probe the mechanisms by which loss of NDH2 func-ion affects parasite metabolism, we considered various enzymeshat would utilize NAD+ in slender BF. One candidate is the glycineleavage complex. However, the excess thymidine in the stan-ard growth medium would be expected to rescue any detrimentalffects on that pathway [13]. Recently it was shown that mito-hondrial production of acetate is essential in BF T. brucei and canroceed through two pathways that contribute roughly equally:ne from pyruvate (the predominant end product of glycolysis) andhe other from threonine, derived from the medium [14]. Enzymesequired for these routes include two NAD+-dependent enzymes,

yruvate dehydrogenase and threonine 3-dehydrogenase, whichere shown to be synthetically lethal (i.e. ablation of either gene

lone was compatible with viability but simultaneous ablation ofoth genes causes death) [21]. We therefore examined by 1H NMR

al Parasitology 211 (2017) 57–61

spectrometry the amounts of three relevant metabolites pyruvate,acetate and alanine (the two latter are minor products of pyruvatemetabolism) excreted by WT, �ndh2, �nubm, and �nubm ndh2cKO parasites from glucose metabolism. No difference in pyruvateand alanine levels was observed between the various parasite lines(Fig. 2A). However, acetate showed a strong reduction in the ndh2null parasites and in the �nubm ndh2 cKO line when NDH2 wasnot expressed (uninduced condition).

The reduced excretion of acetate raised the possibility thatboth cytosolic and mitochondrial fatty acid synthesis might becompromised after loss of NDH2 [22]. We therefore used gas-chromatography mass spectrometry (GC–MS) to quantitate relativechanges in total cellular fatty acid content in the various cell linesafter hydrolyzation of lipid extracts and conversion of the free fattyacids to the corresponding fatty acid methyl esters (FAME) [23]. Forthe �nubm cells the FAME analysis (Fig. 2B) indicated a relativeincrease in C22:4 fatty acids and a relative decrease in C16:0 fattyacids; according to our earlier study, this does not affect growtheither in vitro or in vivo [12]. The most obvious changes in �ndh2parasites were a relative increase in C20:4 fatty acids and, similar to�nubm cells, a relative decrease in C16:0 fatty acids. NDH2 knock-down in the �nubm background also resulted in a reproducible,but for the most part temporary, shift in fatty acid composition thatwas remarkably similar for clones 4 and 5. Forty-eight hours afterwashing away Tet, C18:0 was relatively increased at the expenseof most other fatty acids. After 72 h fatty acid content was largelynormal again for both clones, although clone 5 continued to showa substantial relative reduction for some fatty acids, in particularC20:4.

These findings raised the question as to why the uninduced�nubm ndh2 cKO parasites showed a more severe growth defectand shift in fatty acid content compared to �ndh2 cells, despitea similar reduction in the amount of acetate excreted. There areseveral potential explanations for these observations. It is possiblethat the �ndh2 parasites produced more acetate than uninduced�nubm ndh2 cKO cells, allowing for more de novo fatty acid biosyn-thesis and/or elongation and thus faster growth rates (for technicalreasons we could only measure excretion, not production). How-ever, this explanation seems unlikely given that T. brucei requiresonly a very minor fraction of the acetate it produces for lipidbiosynthesis (∼4% in procyclic forms [14]). Alternatively, some re-expression of NDH2 might have occurred in the �nubm ndh2 cKOcells due to partial loss of repression at the time of the exper-iment (see, for example, Fig. S2). Finally, the growth phenotypeand perturbed fatty acid content may not have been primarily dueto acetate depletion but due to other metabolic pathways thathad been affected by impaired NAD+ regeneration. For example,ablation of the enzyme succinyl-CoA synthetase, activity of whichdepends on acetyl-CoA production and thus secondarily on NAD+

regeneration, was reported to result in the rapid death of BF T. brucei[24].

Taken together, our data shows that NDH2 is an important, butnot essential, factor in maintaining the mitochondrial redox bal-ance in slender BF T. brucei. The temporary cessation of growthof the conditional knockdowns in a genetic background lacking cIfunction, and our inability to obtain mutants that were geneticallynull for both activities, could indicate synthetic lethality of the twoNADH:ubiquinone oxidoreductases. This requires further investi-gation, perhaps with the help of alternative genetic tools such asCre-lox [25] or CRISPR/Cas9 [26–28].

Acknowledgements

Funding: This work was supported by a grant from the NationalInstitutes of Health (USA) [AI 5R01 AI069057] to MP and AS;

Page 5: Molecular & Biochemical Parasitologyschnauferlab.bio.ed.ac.uk/PDFs/Surve et al MBP 2017.pdf · Molecular & Biochemical Parasitology 211 (2017) 57–61 Contents lists available at

hemic

aA(dadaF

A

t1

R

[

[

[

[

[

[

[

[

[

[

[

[

[

[

[

[

[

[Leishmania donovani, MBio 6 (2015) e00861.

S.V. Surve et al. / Molecular & Bioc

grant from the Medical Research Council (UK) [G0600129] toS; grants from the Centre National de la Recherche Scientifique

CNRS, France), The Université de Bordeaux, The Agence Nationalee la Recherche (ANR) [ACETOTRYP of the ANR-BLANC-2010 callnd GLYCONOV of the “Générique” call] and the Laboratoire’Excellence (LabEx) ParaFrap [grant ANR-11-LABX-0024] to FB;nd a grant from the Wellcome Trust [093228] to TKS. We thankred Opperdoes and Paul Michels for very helpful discussions.

ppendix A. Supplementary data

Supplementary data associated with this article can be found, inhe online version, at http://dx.doi.org/10.1016/j.molbiopara.2016.0.001.

eferences

[1] S. Papa, P.L. Martino, G. Capitanio, A. Gaballo, D. De Rasmo, A. Signorile, V.Petruzzella, The oxidative phosphorylation system in mammalianmitochondria, Adv. Exp. Med. Biol. 942 (2012) 3–37.

[2] P. Stenmark, P. Nordlund, A prokaryotic alternative oxidase present in thebacterium Novosphingobium aromaticivorans, FEBS Lett. 552 (2003) 189–192.

[3] A.E. McDonald, S. Amirsadeghi, G.C. Vanlerberghe, Prokaryotic orthologues ofmitochondrial alternative oxidase and plastid terminal oxidase, Plant Mol.Biol. 53 (2003) 865–876.

[4] M. Chaudhuri, R.D. Ott, G.C. Hill, Trypanosome alternative oxidase: frommolecule to function, Trends Parasitol. 22 (2006) 484–491.

[5] A.M.P. Melo, T.M. Bandeiras, M. Teixeira, New insights into type IINAD(P)H:quinone oxidoreductases, Microbiol. Mol. Biol. Rev. 68 (2004)603–616.

[6] F.R. Opperdoes, Compartmentation of carbohydrate metabolism intrypanosomes, Annu. Rev. Microbiol. 41 (1987) 127–151.

[7] B.S.V. Brown, T.B. Chi, N. Williams, The Trypanosoma brucei mitochondrial ATPsynthase is developmentally regulated at the level of transcript stability, Mol.Biochem. Parasitol. 115 (2001) 177–187.

[8] A.G.M. Tielens, J.J. van Hellemond, Surprising variety in energy metabolismwithin Trypanosomatidae, Trends Parasitol. 25 (2009) 482–490.

[9] K. Gunasekera, D. Wüthrich, S. Braga-Lagache, M. Heller, T. Ochsenreiter,Proteome remodelling during development from blood to insect-formTrypanosoma brucei quantified by SILAC and mass spectrometry, BMCGenomics 13 (2012) 556.

10] J.W. Priest, S.L. Hajduk, Developmental regulation of mitochondrial biogenesisin Trypanosoma brucei, J. Bioenerg. Biomembr. 26 (1994) 179–192.

11] E.J. Bienen, M. Saric, G. Pollakis, R.W. Grady, A.B. Clarkson, Mitochondrial

development in Trypanosoma brucei brucei transitional bloodstream forms,Mol. Biochem. Parasitol. 45 (1991) 185–192.

12] S. Surve, M. Heestand, B. Panicucci, A. Schnaufer, M. Parsons, Enigmaticpresence of mitochondrial complex I in Trypanosoma brucei bloodstreamforms, Eukaryot. Cell 11 (2012) 183–193.

[

al Parasitology 211 (2017) 57–61 61

13] A. Roldán, M. Comini, M. a Crispo, R.L. Krauth-Siegel, Lipoamidedehydrogenase is essential for both bloodstream and procyclic Trypanosomabrucei, Mol. Microbiol. 81 (2011) 623–639.

14] M. Mazet, P. Morand, M. Biran, G. Bouyssou, P. Courtois, S. Daulouède, Y.Millerioux, J.-M. Franconi, P. Vincendeau, P. Moreau, Revisiting the centralmetabolism of the bloodstream forms of Trypanosoma brucei: production ofacetate in the mitochondrion is essential for parasite viability, PLoS Negl.Trop. Dis. 7 (2013) e2587.

15] J. Fang, D.S. Beattie, Novel FMN-containing rotenone-insensitive NADHdehydrogenase from Trypanosoma brucei mitochondria: isolation andcharacterization, Biochemistry 41 (2002) 3065–3072.

16] J. Fang, D.S. Beattie, Identification of a gene encoding a 54 kDa alternativeNADH dehydrogenase in Trypanosoma brucei, Mol. Biochem. Parasitol. 127(2003) 73–77.

17] Z. Verner, I. Skodová, S. Poláková, V. Durisová-Benkovicová, A. Horváth, J.Lukes, Alternative NADH dehydrogenase (NDH2):intermembrane-space-facing counterpart of mitochondrial complex I in theprocyclic Trypanosoma brucei, Parasitology 140 (2013) 328–337.

18] M.D. Urbaniak, M.L.S. Guther, M.a.J. Ferguson, Comparative SILAC proteomicanalysis of Trypanosoma brucei bloodstream and procyclic lifecycle stages,PLoS One 7 (2012) e36619.

19] F. Butter, F. Bucerius, M. Michel, Z. Cicova, M. Mann, C.J. Janzen, Comparativeproteomics of two life cycle stages of stable isotope-labeled Trypanosomabrucei reveals novel components of the parasite’s host adaptation machinery,Mol. Cell. Proteomics 12 (2013) 172–179.

20] E. Wirtz, S. Leal, C. Ochatt, G.A.M. Cross, A tightly regulated inducibleexpression system for conditional gene knock-outs and dominant-negativegenetics in Trypanosoma brucei, Mol. Biochem. Parasitol. 99 (1999) 89–101.

21] Y. Millerioux, C. Ebikeme, M. Biran, P. Morand, G. Bouyssou, I.M. Vincent, M.Mazet, L. Riviere, J.-M. Franconi, R.J.S. Burchmore, The threonine degradationpathway of the Trypanosoma brucei procyclic form: the main carbon sourcefor lipid biosynthesis is under metabolic control, Mol. Microbiol. 90 (2013)114–129.

22] S.H. Lee, J.L. Stephens, P.T. Englund, A fatty-acid synthesis mechanismspecialized for parasitism, Nat. Rev. Microbiol. 5 (2007) 287–297.

23] S.O. Oyola, K.J. Evans, T.K. Smith, B.A. Smith, J.D. Hilley, J.C. Mottram, P.M.Kaye, D.F. Smith, Functional analysis of Leishmania cyclopropane fatty acidsynthetase, PLoS One 7 (2012) e51300.

24] X. Zhang, J. Cui, D. Nilsson, K. Gunasekera, A. Chanfon, X. Song, H. Wang, Y. Xu,T. Ochsenreiter, The Trypanosoma brucei MitoCarta and its regulation andsplicing pattern during development, Nucleic Acids Res. 38 (2010) 7378–7387.

25] H.-S. Kim, Z. Li, C. Boothroyd, G.a.M. Cross, Strategies to construct null andconditional null Trypanosoma brucei mutants using Cre-recombinase andloxP, Mol. Biochem. Parasitol. 191 (2013) 16–19.

26] N. Lander, Z.-H. Li, S. Niyogi, R. Docampo, CRISPR/Cas9-induced disruption ofparaflagellar rod protein 1 and 2 genes in Trypanosoma cruzi reveals their rolein flagellar attachment, MBio 6 (2015) e01012.

27] W.-W. Zhang, G. Matlashewski, CRISPR-Cas9-mediated genome editing in

28] L. Sollelis, M. Ghorbal, C.R. MacPherson, R.M. Martins, N. Kuk, L. Crobu, P.Bastien, A. Scherf, J.-J. Lopez-Rubio, Y. Sterkers, First efficientCRISPR-Cas9-mediated genome editing in Leishmania parasites, Cell.Microbiol. 17 (2015) 1405–1412.

Page 6: Molecular & Biochemical Parasitologyschnauferlab.bio.ed.ac.uk/PDFs/Surve et al MBP 2017.pdf · Molecular & Biochemical Parasitology 211 (2017) 57–61 Contents lists available at

Supplementary Material

Supplemental Methods

Transfection and Growth assays. T. brucei bloodstream forms (BF) were grown in HMI-9

medium with 10% heat-inactivated fetal bovine serum at 37°C and 5% CO2. Drugs to maintain

selection for transgenes were included when relevant: hygromycin (1 g/ml), puromycin (0.1

g/ml), blasticidin S (3 μg/ml) and phleomycin (1 μg/ml). BF were transfected using Amaxa

Nucleofector program X-001 [1,2]. Briefly, 3X107 log phase BF cells were harvested by

centrifugation and pellets were resuspended in 100 μl of Human T-cell Nucleofector solution

plus 2-3 μg of DNA construct. Transfected cells were diluted 1:10 and 1:100 and incubated at

37°C for at least 6 h before adding appropriate drugs as described above. Protein expression

from tetracycline (Tet)-regulated constructs was induced with 1 to 2 g/ml of Tet. In vitro and in

vivo growth analysis was carried as described [3].

Conditional ndh2 knockouts in Lister 427 BF. Primers are listed in Table S1. Deletion constructs

were generated using plasmids pLew13 and pLew90 [4]. The regions targeting the drug

resistance cassette were 500 bp region upstream of the NDH2 coding sequence (CDS),

Tb927.10.9440 (amplified using PCR primers 1 and 2 with NotI and MluI sites, respectively) and

the last 456 bp from the 3’ end of the NDH2 CDS (amplified using primers 3 and 4, with XbaI

and StuI sites, respectively). These regions were used to flank T7 RNA polymerase and NEO

genes (first allele knockout) and the Tet repressor and HYG gene cassette (second allele

knockout), as described [5]. For ectopic expression, V5-tagged NDH2 was generated by cloning

the NDH2 CDS into the pLew-3V5-PAC plasmid [3] via HindIII and BamHI sites (primers 17 and

18). Constructs were confirmed by restriction analysis and sequencing, and were digested with

NotI before transfection.

Null mutants in single marker BF and ndh2 cKOs in Δnubm BF. A stitching PCR approach was

employed to generate deletion constructs [6]. The constructs were designed such that drug

resistance genes were flanked by regions upstream and the last 456 bp of the NDH2 CDS.

First, these regions were amplified from genomic DNA using Phusion polymerase system

(Thermo) with an overhang (20 bp) of the sequence from drug resistance genes (HYG, PHLEO

or PAC) (primers 5-10). CDSs for the drug resistance genes were also amplified using Phusion

PCR system (primers 13-16). The first round amplicons were then stitched together in by

amplification using primers situated at 5’ end of the 5’ UTR and 3’ end of the 3’UTR (primers 11

and 12). The second round amplicons were then dA-tailed using GoTaq Polymerase (Promega),

and cloned into pGEM-T Easy (Promega) vector. An ectopic copy of V5 tagged NDH2 was

generated by cloning the NDH2 CDS into pLew79-3V5-BSD, which was constructed by

swapping the PAC gene from pLew79-NDH2-V5-PAC plasmid with the BSD gene from

pLew100-V5-BSD (kind gift from George Cross) using SnaBI sites.

Initial confirmation of gene disruptions was done by PCR. Subsequent Southern analysis

confirmation was performed as shown in the main text.

Page 7: Molecular & Biochemical Parasitologyschnauferlab.bio.ed.ac.uk/PDFs/Surve et al MBP 2017.pdf · Molecular & Biochemical Parasitology 211 (2017) 57–61 Contents lists available at

Table S1. Primers

Primer Name Sequence (5’-3’) Use

Primers for NDH2 cKO 1 Ndh 5F ATGAGCGGCCGCATTTTATGTGTGAAGGGGAGG Amplify 5’ UTR 2 Ndh 5R ATGAACGCGTCTCGAGACTTGTCGTTCAGATCTCCACT Amplify 5’ UTR 3 Ndh 3F ATATCTAGAATTTAAATCTGCTGTCGCTTCTCGTCA Amplify 3’ CDS end 4 Ndh 3R ATAAGGCCTGCGGCCGCCTACTCGTTCTGTTTCTTTG Amplify 3’ CDS end

Primers for stitching KOs 5 Ndh 5F-642 AGATATTGTCCGTACTGCCTC fusion 5’ UTR 6 NDH 5R Phleo GAACGGCACTGGTCAACTTGGCCATCTTCCTTCTTTTTCCTCCCCTC Fusion 5’ UTR to PHLEO 7 NDH 5R PAC GTGGGCTTGTACTCGGTAACCATCTTCCTTCTTTTTCCTCCCCTC Fusion 5’UTR to PAC 8 NDH 3F Phleo GTGGCCGAGGAGCAGGACTGAGTGAAGTTTGTGTACCAAGCAAC Fusion 3’ UTR to PHLEO 9 NDH 3F PAC GACCCGCAAGCCCGGTGCCTGAGTGAAGTTTGTGTACCAAGCAAC Fusion 3’ UTR to PAC

10 Ndh 3R+931 AGAGGAGGCAAAAAGGTTCA Fusion 3’ UTR 11 Ndh 5F-566 TGTGTGGAAAGGGAACGAAG Final fusion PCR 12 Ndh 3R+856 GAGAGTGCGCCAGTGTCATAC Final fusion PCR 13 Phleo F ATG ATGGCCAAGTTGACCAGTGCCGTTC Amplify resistance gene 14 Phleo R Stop TCAGTCCTGCTCCTCGGCCAC Amplify resistance gene 15 PAC F ATG ATGGTTACCGAGTACAAGCCCAC Amplify resistance gene 16 PAC R Stop TCAGGCACCGGGCTTGCGGGTC Amplify resistance gene

Primers for complementation 17 Ndh2 Hind F ATGCAAGCTTATGATATGCCGCACATCTT Cloning Gene 18 Ndh2 Bam R GCATGGATCCCTCGTTCTGTTTCTTTGTCG Cloning Gene

Restriction sites added for cloning are underlined. Bases in bold anneal with the drug resistance genes in the fusion

PCR, while standard font matches NDH2 flanking sequence. In the primer names, negative numbers indicate nt

upstream of NDH2 start codon (forward) and positive number indicate position downstream of NDH2 stop codon

(reverse primers).

References

[1] Burkard G, Fragoso CM, Roditi I. (2007) Highly efficient stable transformation of bloodstream forms

of Trypanosoma brucei. Mol Biochem Parasitol 153: 220-3.

[2] Cross GAM. Tools for genetic analysis in Trypanosoma brucei. Available from URL:

http://tryps.rockefeller.edu/trypsru2_protocols_index.html

[3] Surve S, Heestand M, Panicucci B, Schnaufer A, Parsons M. (2011) Enigmatic presence of

mitochondrial complex I in Trypanosoma brucei bloodstream forms. Eukaryot Cell 11: 183-93.

[4] Wirtz E, Leal S, Ochatt C, Cross GA. (1999) A tightly regulated inducible expression system for

conditional gene knock-outs and dominant-negative genetics in Trypanosoma brucei. Mol Biochem

Parasitol 99: 89-101.

[5] Ochatt CM, Butikofer P, Navarro M, Wirtz E, Boschung M, Armah D, et al. (1999) Conditional

expression of glycosylphosphatidylinositol phospholipase C in Trypanosoma brucei. Mol Biochem

Parasitol 103: 35-48.

[6] Merritt C, Stuart K. (2013) Identification of essential and non-essential protein kinases by a fusion

PCR method for efficient production of transgenic Trypanosoma brucei. Mol Biochem Parasitol

190: 44-9.

Page 8: Molecular & Biochemical Parasitologyschnauferlab.bio.ed.ac.uk/PDFs/Surve et al MBP 2017.pdf · Molecular & Biochemical Parasitology 211 (2017) 57–61 Contents lists available at

NDH2S S

CDS+flankCDS 1 kb

AWT a1-4 a3-10

2

4

6

8

1012

ndh2 ndh2WT a1-4 a3-10

CDS CDS+flankProbe

B

1.E+05

1.E+06

1.E+07

1.E+08

1.E+09

1.E+10

1.E+11

0 1 2 3 4 5

Cu

mm

ual

tive

ce

lls/m

l

Day

WT1

WT2

SKO a1

DKO a1-4

SKO a3

DKO a3-10

1010

108

107

106

105

109

40

60

50

70

anti-V5

anti-PGK

kDa0 72 0 2 7

clone 4 clone 5

days

50

60

Fig. S1. Generation and growth analysis of ndh2 cell lines. (A) Genomic Southern blot of ∆ndh2clones. Genomic DNA was digested with SnaBI (S), fractionated by agarose gel electrophoresis,transferred to a membrane and probed with a fragment consisting of either the coding sequence (CDS)alone or the CDS plus 5’ and 3’ flanking regions (CDS+flank) as shown below. Expected sizes for the CDSprobe are: WT, 3.6 kbp; ∆ndh2 clones a1-4 and a3-10, no band. For the ‘CDS+flank’ probe expectedsizes are: WT, 3.6 and 4.1 kbp; ∆ndh2 clones, 2.5, 2.7, and 4.1 kbp. (B) Growth curve of ∆ndh2 (DKO)mutants compared to their parental single knockout clones (SKO) and the parental single marker line(WT). The curve shows the cumulative cell concentration per ml with error bars marking the standarddeviation of the triplicate data points. This experiment was conducted at 46 days of culture post-transfection.

Fig. S2. Re-expression of NDH2-V5 in the ∆nubm ndh2cKOs several days after Tet withdrawal. Western blotanalysis showing that after withdrawal of Tet, NDH2-V5expression had strongly decreased by day 2 as comparedto day 0, but increased again at later time points. Thesame blot was re-probed with anti-PGK as loadingcontrol. The samples analyzed here were derived from arepeat of the growth experiment shown in Fig. 1E.Proliferation slowed from day 1 to day 3 and thenrecovered to near normal rates by day 5.

Supplementary Figures