bioRxiv preprint doi: https://doi.org/10.1101/2021.04.20.440500; this version posted April 21, 2021. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY-NC-ND 4.0 International license.

1 TCAIM controls effector generation by preventing Mitochondria-Endoplasmic 2 Reticulum Contact Site-initiated Cholesterol Biosynthesis 3 4 Christina Iwert1, Julia Stein1, Christine Appelt1, Katrin Vogt1, Roman Josef Rainer2, Katja 5 Tummler2, Kerstin Mühle1, Katarina Stanko1, Julia Schumann1, Doreen Uebe1, Karsten 6 Jürchott3, Jan Lisec4, Katharina Janek5, Christoph Gille5, Kathrin Textoris-Taube6, Somesh 7 Sai7, Ansgar Petersen3, Anja A. Kühl8, Edda Klipp2, Christian Meisel1, Birgit Sawitzki1* 8

9 1 Charité – Universitätsmedizin Berlin, corporate member of Freie Universität Berlin and 10 Humboldt-Universität zu Berlin, Institute of Medical Immunology, Germany 11 2 Humboldt-Universität zu Berlin, Institute of Biology, Theoretical Biophysics, Germany

12 3 Berlin Institute of Health at Charité – Universitätsmedizin Berlin, BIH Center for 13 Regenerative Therapies BCRT, Germany 14 4 Bundesanstalt für Materialforschung und -prüfung (BAM), Department of Analytical 15 Chemistry and Reference Materials, Germany

16 5 Charité – Universitätsmedizin Berlin, corporate member of Freie Universität Berlin and 17 Humboldt Universität zu Berlin, Institute for Biochemistry, Germany

18 6 Charité – Universitätsmedizin Berlin, corporate member of Freie Universität Berlin and 19 Humboldt-Universität zu Berlin, Institute for Biochemistry, Core Facility High-Throughput 20 Mass Spectrometry, Germany 21 7 Max Delbrueck Center for Molecular Medicine Berlin, Scientific Genomics Platform, 22 Germany

23 8 Charité – Universitätsmedizin Berlin, corporate member of Freie Universität Berlin and 24 Humboldt-Universität zu Berlin, iPATH.Berlin – Core Unit Immunopathology for Experimental 25 Models, Germany 26 * Corresponding author (e-mail: [email protected]) 27

1

bioRxiv preprint doi: https://doi.org/10.1101/2021.04.20.440500; this version posted April 21, 2021. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY-NC-ND 4.0 International license.

28 Abstract 29 T cells need to adapt their cellular metabolism for effector cell differentiation. This relies on 30 alterations in mitochondrial physiology. Which signals and molecules regulate those 31 alterations remains unclear. We recently reported, that the mitochondrial TCAIM 32 inhibits activation-induced changes in mitochondrial morphology and function and thus, CD4+ 33 effector T cell formation. Using conditional TCAIM knock-in (KI) and knockout (KO) mice, we 34 now show that it also applies to CD8+ T cells and more importantly, delineate the molecular 35 processes in mitochondria by which TCAIM controls effector cell differentiation. TCAIM KI 36 resulted in reduced activation-induced HIF1α protein expression. Metabolomics and 37 transcriptional data in combination with mathematical flux modeling revealed an impaired 38 induction of anabolic pathways, especially of the mevalonate pathway and cholesterol 39 biosynthesis in TCAIM KI CD8+ T cells. Addition of cholesterol completely rescued HIF1α 40 protein expression, activation and proliferation of TCAIM KI CD8+ T cells. At the molecular 41 level, TCAIM delayed activation-induced mitochondria-ER contact (MERC) formation by 42 binding to MERC promoting such as RMD3 and VDAC2. In summary, we 43 demonstrate that TCAIM suppresses effector cell differentiation by inhibiting MERC 44 formation, which induce HIF1α-mediated increase in cellular metabolism and cholesterol 45 biosynthesis. 46

2

bioRxiv preprint doi: https://doi.org/10.1101/2021.04.20.440500; this version posted April 21, 2021. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY-NC-ND 4.0 International license.

47 Introduction 48 In order to cope with increased metabolic demands arising with T cell activation and effector 49 differentiation, T cells need to upregulate catabolic but also to induce anabolic processes. 50 Upregulation of catabolic processes mainly comprises elevations of aerobic glycolysis, 51 glutaminolysis and oxidative (OXPHOS) necessary for profound building 52 block(s) and adenosine triphosphate (ATP) synthesis. Newly generated building blocks are 53 then funneled into anabolic pathways to enable nucleotide, protein and lipid synthesis to 54 ensure DNA replication, cell growth and proliferation as well as the acquisition of effector 55 functions1. 56 Especially, effector T cell differentiation and function relies on upregulation of aerobic 57 glycolysis2, which is regulated by the two transcription factors MYC and HIF1α. Whereas 58 MYC is essential to initiate aerobic glycolysis during T cell activation3, HIF1α sustains aerobic 59 glycolysis throughout T cell activation to enforce effector T cell differentiation and, more 60 importantly, effector function4. HIF1α signaling has been shown to be indispensable for 61 Granzyme B expression, CD62L downregulation and CD8+ T cell migration4, 5. Therefore, 62 HIF1α activation is crucial for effector T cell function and recruitment to the inflammatory 63 environment. 64 Metabolic adaptations necessary for effector T cell differentiation have been shown to 65 depend on alterations in mitochondrial physiology6. OXPHOS is strongly upregulated upon 66 T cell activation, which is accompanied by an increase in mitochondrial membrane potential 67 (ΔΨm) and mitochondrial reactive oxygen species (mROS) generation7. Whereas an 68 increase in ΔΨm is associated with an enhanced effector differentiation and, more 69 importantly, effector cytokine secretion in T cells8, mROS have been shown to act as redox 70 signaling molecules regulating expression and proliferation9, 10, 11. Besides mROS, 71 mitochondria also actively shape cellular transcription through the release of TCA cycle 72 derived metabolites12. Moreover, effector T cell differentiation has been linked to 73 mitochondrial fission and cristae resolution as a prerequisite for acquisition of aerobic 74 glycolysis13. Failure to engage either will result in deregulated T cell activation and 75 differentiation14. Thus, mitochondria function as central immunometabolic hubs to guide T cell 76 activation and differentiation. However, which signals and molecules regulate those 77 activation-induced alterations in mitochondrial physiology still remains largely unclear. 78 We recently reported that the mitochondrial protein T cell activation inhibitor, mitochondrial 79 (TCAIM) is downregulated upon T cell activation. Notably, reinforced TCAIM expression 80 inhibits effector T cell differentiation of conventional CD4+ T cells leading to full acceptance of 81 allogeneic skin grafts15. TCAIM exclusively locates within mitochondria16 where it might 82 interfere with activation-induced changes in mitochondrial morphology and function15. 83 Here, we hypothesized that TCAIM acts as an important link between activation-induced 84 mitochondrial alterations and changes in cellular metabolism driving full effector cell 85 differentiation. To investigate this, we utilized conditional TCAIM KI and KO mice, 86 overexpressing TCAIM or displaying a loss of function in the TCAIM protein, to examine 87 which molecular processes in mitochondria control CD8+ T cell effector differentiation. We 88 found that TCAIM KI strongly impaired CD8+ T cell proliferation and effector differentiation as 89 well as upregulation of HIF1α protein expression and transcription of encoding for 90 enzymes of the mevalonate pathway and cholesterol biosynthesis. Importantly, addition of 91 cholesterol rescued T cell activation and proliferation of TCAIM KI CD8+ T cells. At the 92 molecular level, we showed that TCAIM delays activation-induced interaction between 93 mitochondria and the endoplasmic reticulum (ER) by binding to voltage-dependent 94 anion-selective channel 2 (VDAC2) and regulator of microtubule dynamics protein 3 (RMD3), 95 both of which promote mitochondria-ER contact (MERC) formation. In summary, we provide 3

bioRxiv preprint doi: https://doi.org/10.1101/2021.04.20.440500; this version posted April 21, 2021. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY-NC-ND 4.0 International license.

96 evidence that TCAIM negatively regulates the activation-induced increase in MERC 97 formation, thereby, preventing the upregulation of anabolic processes, particularly the 98 mevalonate pathway and cholesterol biosynthesis, thus, resulting in impaired CD8+ T cell 99 activation and proliferation. 100

4

bioRxiv preprint doi: https://doi.org/10.1101/2021.04.20.440500; this version posted April 21, 2021. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY-NC-ND 4.0 International license.

101 Results 102 TCAIM KI CD8+ T cells show reduced glycolytic flux and biosynthesis as well as 103 impaired CD8+ T cell differentiation 104 Upregulation of anabolic processes is a hallmark of T cell activation and differentiation. To 105 investigate whether TCAIM interferes with CD8+ T cell anabolic upregulation upon activation, 106 we performed a non-targeted metabolite analysis. For this, we used mice with a conditional 107 T cell specific Tcaim gene overexpression (KI) and Cd4-Cre wild type (wt) controls from 108 littermates or parallel breeding. Isolated naïve CD8+ T cells from spleen and lymph nodes of 109 these mice were polyclonally activated in vitro with αCD3 and αCD28 antibodies for 60 h. 110 Metabolite levels were measured using gas chromatography mass spectrometry 111 (GC/APCI-MS). 112 Comparing the levels of a chosen fraction of detected metabolites from TCAIM KI CD8+ 113 T cells with that of wt CD8+ T cells we found, that TCAIM KI CD8+ T cells showed increased 114 glucose levels, while intermediates and especially end products of the glycolytic pathway like 115 pyruvate or lactic acid were reduced suggesting a diminished glycolytic flux due to TCAIM KI 116 (Fig. 1a upper row and Extended Data Fig. 1a). Also, intermediates of the phospholipid 117 biosynthesis pathway like glycerol-3-phosphate or sphingosine as well as levels of 118 non-essential amino acids derived either from glycolysis or TCA cycle intermediates e.g. 119 alanine and or proline and aspartate, respectively, were reduced (Fig. 1a middle and 120 lower row). In contrast, citrate and α-ketoglutarate early intermediates of the TCA cycle were 121 increased while late TCA cycle intermediates like malate or fumarate were strongly 122 decreased (Extended Data Fig. 1b). Additionally, glutamine, starting point of the 123 glutaminolysis pathway, was enriched while glutamate showed similar levels (Extended Data 124 Fig. 1c) suggesting a reduced support of the TCA cycle through a diminished glutamine 125 breakdown. Interestingly, although reduced glycolysis and TCA cycle end products or lower 126 levels of biosynthesis related metabolites suggest an overall diminished metabolism in 127 TCAIM KI CD8+ T cells, intermediates of the pentose phosphate pathway as 128 ribose-5-phosphate and other sugar metabolites such as fructose were enriched in TCAIM KI 129 CD8+ T cells (Extended Data Fig. 1d). 130 Taken together, the non-targeted metabolite analysis revealed an altered metabolic program 131 in activated TCAIM KI CD8+ T cells with reduced metabolite levels of glycolytic end products 132 and biosynthesis pathways when compared to wt CD8+ T cells. 133 As a high glycolytic flux is essential for effector T cell differentiation, we next investigated 134 whether TCAIM KI interferes with CD8+ T cell activation and differentiation. Again, naive 135 TCAIM KI and wt CD8+ T cells were polyclonally activated and changes in CD62L and CD44 136 surface expression were measured by flow cytometry to discriminate between naïve 137 (CD62L+ / CD44–), central memory (CD62L+ / CD44+) and effector/ effector memory 138 (CD62L– / CD44+) T cells. We additionally included CD8+ T cells isolated from mice with a 139 conditional T cell specific knock-out (KO) of exon 4 in the Tcaim gene, which results in a loss 140 of protein function, as well as strain specific Cd4-Cre wt controls to the analysis. 141 After 72 h of activation the majority of TCAIM KI CD8+ T cells maintained a naïve phenotype 142 expressing CD62L only. Furthermore, the few activated TCAIM KI CD8+ T cells rather 143 obtained a central memory phenotype expressing both CD62L and CD44. In contrast, TCAIM 144 KO or wt CD8+ T cells were predominantly activated and differentiated into either effector/ 145 effector memory or central memory cells (Fig. 1b,c). 146 We also analyzed expression of the two activation markers CD69 and CD25 indicative of the 147 tissue homing and proliferative potential, respectively. Indeed, TCAIM KI also prevented 148 upregulation of CD69 (Fig. 1d) and CD25 (Fig. 1e) expression in comparison to wt and 149 TCAIM KO CD8+ T cells. However, CD69 and CD25 expression was not completely inhibited 150 in TCAIM KI CD8+ T cells indicating residual T cell activation capacity. Nevertheless, this 5

bioRxiv preprint doi: https://doi.org/10.1101/2021.04.20.440500; this version posted April 21, 2021. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY-NC-ND 4.0 International license.

151 residual activation was not sufficient to acquire full proliferative potential (Fig. 1f,g) nor 152 effector function (Fig. 1h,i). 153 Since TCAIM KI impairs CD8+ T cell activation and restricts differentiation, we expected an 154 increased CD8+ T cell activation, proliferation and especially effector/ effector memory 155 differentiation and function due to TCAIM KO. Yet, no significant differences between TCAIM 156 KO and wt CD8+ T cells upon in vitro activation could be observed (Fig. 1b-i). However, ex 157 vivo characterization of TCAIM KO CD8+ T cells by flow cytometry revealed a tendency 158 towards increased spontaneous activation and effector/ effector memory differentiation of 159 TCAIM KO CD8+ T cells in the colon (Extended Data Fig. 2a,b). Additionally, promoted a 160 homozygous (hocre) TCAIM KO an age-dependent accumulation of effector/ effector 161 memory CD8+ T cells in the colon, which could not be observed for heterozygous (hecre) or 162 wild type (wtcre) TCAIM KO mice (Extended Data Fig. 2c). Furthermore, TCAIM KO mice 163 showed an increased intra-epithelial infiltration rate of CD8+ T cells in the colon (Fig. 1j,k). 164 We did not detect differences in frequencies and absolute numbers of total CD8+ T cells in 165 lymphatic or non-lymphatic tissues, e.g. colon, from TCAIM KO mice (Extended Data 166 Fig. 2d,e). 167 Collectively, the data indicate that TCAIM KI leads to an altered metabolic program with 168 reduced biosynthesis processes and suppressed CD8+ T cell activation and especially 169 effector/ effector memory differentiation. In contrast, a TCAIM KO seems to be advantageous 170 for CD8+ T cell activation, proliferation and intra-tissue accumulation of effector/ effector 171 memory CD8+ T cells. 172 173 Tcaim is regulated by NFAT 174 Knowing that TCAIM interferes with effector T cell differentiation and that its expression is 175 downregulated upon activation15, we investigated the transcriptional control of its promoter. 176 The promotor region of the murine Tcaim gene has predicted binding sites for nuclear factor 177 of activated T cells (NFAT), CCAAT/enhancer-binding-proteine (C/EBP) and cAMP response 178 element-binding protein (CREB) (Fig. 2a). Especially, NFAT function is essential for full T cell 179 activation and acquisition of effector functions17. NFAT is activated by dephosphorylation 180 through calcineurin18. Thus, we hypothesized NFAT as a negative regulator for Tcaim gene 181 expression. To test this, we added the calcineurin inhibitor cyclosporine A (CsA) at 182 increasing doses to stimulation cultures of naïve CD8+ T cells isolated from C57BL6/N mice. 183 Tcaim mRNA expression was measured by real-time qPCR. 184 DMSO control-treated CD8+ T cells showed a significant downregulation of Tcaim mRNA 185 expression over time while CsA treatment resulted in a dose-dependent inhibition of Tcaim 186 mRNA downregulation at 48 h of activation (Fig. 2b). Effective NFAT inhibition through CsA 187 additionally was controlled by measuring mRNA expression of Il-2 and Ifn-γ mRNA as two 188 main target genes of NFAT. Expression of both was significantly reduced with increasing 189 doses of CsA, while untreated cells showed an increase in mRNA expression levels upon 190 activation (Fig. 2c). 191 Taken together, NFAT indeed acts as a negative regulator for Tcaim gene expression. 192 193 Metabolic modeling predicts forward directed glycolytic TCA cycle activity in general 194 and reduced metabolic flux in TCAIM KI CD8+ T cells in particular 195 To further dissect the consequences of TCAIM KI for activation-induced upregulation of 196 anabolic processes in CD8+ T cells, we analyzed the gene expression profile of naïve and 197 polyclonally activated CD8+ T cells 24 to 48 h after stimulation from TCAIM KI, KO or wt mice

6

bioRxiv preprint doi: https://doi.org/10.1101/2021.04.20.440500; this version posted April 21, 2021. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY-NC-ND 4.0 International license.

198 by RNA-seq. 199 A targeted gene expression analysis of selected genes encoding for metabolically relevant 200 transcription factors and enzymes of the glycolysis, tricarboxylic acid (TCA) cycle and 201 glutaminolysis pathway revealed mainly a diminished upregulation of genes encoding for 202 enzymes or enzyme subunits of the TCA cycle or glutaminolysis in both 24 and 48 h 203 activated TCAIM KI compared to wt CD8+ T cells (Fig. 3a genes in bold). 204 The two main transcription factors for initiating and sustaining aerobic glycolysis and thus, 205 activation-dependent metabolic adaptations of T cells, are MYC and HIF1α. Gene expression 206 of both were upregulated in TCAIM KI CD8+ T cells and showed levels comparable to that of 207 TCAIM KO and wt CD8+ T cells (Fig. 3a and Extended Data Fig. 3a). In accordance to 208 upregulation to that, gene expression of glycolytic enzymes was upregulated in TCAIM KI, 209 KO and wt CD8+ T cells 48 h after activation to a similar extend. This was unexpected, as we 210 assumed a reduced glycolytic activity represented by a reduced glycolysis gene expression 211 in TCAIM KI CD8+ T cells according to our previous observations of diminished glycolysis 212 end product metabolite levels and abrogated effector differentiation. In contrast, there were 213 genes with even significantly increased expression in TCAIM KI compared to wt CD8+ T cells 214 either due to reduced down- or increased upregulation (e.g. Ldhd, Pgam2 or Ldhb, Pkm, 215 Slc2a1, respectively). Only, Aldoc, Eno3 and the HIF1α target gene Hk2 were less 216 upregulated in activated TCAIM KI CD8+ T cells with Eno3 being the only significant one 217 (Fig. 3a). Differences in metabolic enzyme gene expression between activated TCAIM KO 218 and wt CD8+ T cells could not be observed (Fig. 3a). 219 Thus, in accordance with the metabolite analysis, we observed reduced upregulation of TCA 220 cycle and glutaminolysis but not for the majority of glycolysis controlling genes suggesting at 221 least reduced metabolic rates in mitochondria of activated TCAIM KI CD8+ T cells. 222 The pleiotropic changes in the gene expression patterns cannot be directly linked to changes 223 in metabolic fluxes, as the intermediate steps of protein translation and regulation of enzyme 224 activity were not considered. We therefore mathematically constructed a model, which 225 incorporated the available metabolomics data on TCAIM KI and wt CD8+ T cells to estimate 226 their metabolic fluxes, with the TCAIM KI model being a differentially regulated variant of the 227 wt model. The models shall foster the understanding of the general distribution of metabolic 228 activity in activated CD8+ T cells, but also, the specific effects of TCAIM affecting this 229 well-balanced process. 230 The calculated metabolic model of polyclonally activated wt CD8+ T cells revealed a strong 231 support of the TCA cycle activity by increased glycolysis but also glutaminolysis rates 232 (Fig. 3b grey bar charts of pink and yellow shaded paths). The TCA cycle flux was forward 233 directed with no evidence of reversed citrate generation from glutamine to support biomass 234 production. However, to ensure continuous forward TCA cycle flux, wt CD8+ T cells had to 235 convert malate to pyruvate by malic enzyme (ME) to provide enough pyruvate and 236 acetyl-CoA for increased citrate synthase (CS) activity (Fig. 3b grey ME bar chart). Similar to 237 wt CD8+ T cells, activated TCAIM KI CD8+ T cells also showed an exclusively forward 238 directed TCA cycle flux, which was driven by glycolysis and glutaminolysis (Fig. 3b blue bar 239 charts of pink and yellow shaded paths). However, overall flux rates were strongly diminished 240 in TCAIM KI CD8+ T cells (Fig. 3b blue bar charts). Also, differing to wt CD8+ T cells, 241 glycolysis derived pyruvate was sufficient to meet their calculated CS activity and thus, 242 showed no malate to pyruvate conversion. 243 In summary, the mathematical flux model suggests a continuous glycolytic TCA cycle flux in 244 activated wt CD8+ T cells which was supported by increased glutaminolysis and malate to 245 pyruvate conversion. In contrast, for activated TCAIM KI CD8+ T cells an overall reduced 246 metabolic flux was proposed. However, except for the step of malate to pyruvate conversion 247 directions of the major fluxes were not altered in comparison to wt CD8+ T cells.

7

bioRxiv preprint doi: https://doi.org/10.1101/2021.04.20.440500; this version posted April 21, 2021. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY-NC-ND 4.0 International license.

248 249 Decreased metabolic flux rates of activated TCAIM KI CD8+ T cells align with a reduced 250 enzyme gene expression, ΔΨm and HIF1α protein expression 251 To validate the findings of the mathematically evaluated metabolic fluxes of TCAIM KI and wt 252 CD8+ T cells, we compared gene expression of differentially expressed enzymes between 253 polyclonally activated TCAIM KI and wt CD8+ T cells by RNA-seq (Extended Data Fig. 3b) 254 with a regularization parameter introduced to calculate the deviating flux rates of TCAIM KI 255 from that of wt CD8+ T cells. The regularization parameter describes the alteration of flux 256 through each modelled enzyme summarizing the effects of different concentrations of 257 enzyme substrates and products but also of co-factor availability. A negative regularization 258 parameter value indicates a reduced activity of a particular enzyme and should align with a 259 negative fold change in gene expression which indicates either a downregulation or reduced 260 upregulation of this enzyme in TCAIM KI CD8+ T cells equivalent to a reduced enzyme 261 activity. We thereby assume that mRNA levels correlate, at least to some extent, to enzyme 262 activity. 263 As can be seen in Fig. 4a most of the enzyme regularization parameters aligned with the 264 negative fold changes of the respective enzyme coding genes (Fig. 4a lower left corner). For 265 the alanine transaminase (ALT) enzyme a regularization parameter for the backward flux rate 266 (kb) was calculated, which is equivalent to a downregulated forward flux rate (kf). Thus, also 267 the ALT regularization parameter agreed with a reduced Gpt2 (encoding for ALT) expression 268 in TCAIM KI CD8+ T cells. The mismatches of the other calculated regulation parameters 269 with their respective enzyme gene expressions in the upper left and lower right corners might 270 be explained by missing information on either substrate and/ or product (GAPDH, SUCLG2 271 or PGM) as well as co-factor concentrations (MDH, PDH, GLUD, IDH2, LDH and PK). 272 Especially the lack of knowledge about NAD+/ NADH++H+ ratios in wt and TCAIM KI CD8+ 273 T cells complicates a reliable regularization parameter calculation. However, assuming an 274 increased NAD+/ NADH++H+ ratio in TCAIM KI CD8+ T cells due to a decreased TCA cycle 275 activity would decrease the regularization parameter of MDH, PDH, GLUD and IDH2 and 276 increase the regularization parameter of LDH and, thereby, increase the chance of alignment 277 with the fold changes in enzyme gene expression. Additionally, it should be kept in mind that 278 enzyme gene expression is not perfectly ideal for metabolic flux modelling confirmation as 279 many enzymes are regulated at a post-transcriptional level and may have other regulatory or 280 structural functions such as GAPDH regulating IFN- protein expression2. 281 The integrity of ΔΨm is closely linked to the TCA cycle and OXPHOS activity. NADH+H+ 282 generated by the TCA cycle is used as electron donor by the respiratory complexes19. In 283 consequence of the electron transport, a proton gradient is built up, which is reflected by the 284 ΔΨm and used for ATP synthesis by OXPHOS20. As we observed a reduced TCA cycle 285 activity in the metabolic flux model of TCAIM KI CD8+ T cells, we hypothesized that they also 286 have a lower ΔΨm. We performed flow cytometry analysis with naïve and polyclonally 287 activated TCAIM KI, KO and wt CD8+ T cells and measured the ΔΨm using the fluorescent 288 dye TMRE. 289 All CD8+ T cells responded to the activation stimulus with an increase of the ΔΨm reaching a 290 peak after 72 h, but as expected, the increase was significantly less in TCAIM KI CD8+ 291 T cells (Fig. 4b). Additionally, the baseline ΔΨm was also significantly lower in TCAIM KI 292 CD8+ T cells compared to the respective wt control while it tended to be higher in TCAIM KO 293 CD8+ T cells (Fig. 4c). 294 A prolonged high glycolytic flux in activated T cells is controlled by the transcription factor 295 HIF1α4. To ensure rapid adaptation to changes in oxygen availability or immune stimulation 296 HIF1α is constantly expressed. Under homeostatic conditions in non-activated T cells the 8

bioRxiv preprint doi: https://doi.org/10.1101/2021.04.20.440500; this version posted April 21, 2021. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY-NC-ND 4.0 International license.

297 protein is immediately degraded, but upon T cell activation HIF1α mRNA and protein 298 expression becomes upregulated and HIF1α protein degradation is inhibited21. The latter is 299 especially linked to increased levels of the TCA cycle metabolites succinate22 and fumarate23 300 but also to increased mROS level21. 301 As mentioned before, Hif1a mRNA expression was upregulated upon activation in both 302 TCAIM KI and wt CD8+ T cells. However, as HIF1α is primarily regulated at the 303 post-transcriptional level and the mathematical metabolic modelling of TCAIM KI CD8+ 304 T cells predicted a reduced TCA cycle activity correlating with reduced levels of TCA cycle 305 derived succinate and fumarate, we hypothesized that HIF1α protein expression was 306 reduced upon activation of TCAIM KI CD8+ T cells. Thus, we measured HIF1α protein 307 expression by flow cytometry. Indeed, TCAIM KI CD8+ T cells showed significantly 308 decreased HIF1α protein levels upon activation compared to wt CD8+ T cells in a gene dose 309 depend manner (Fig. 4d,e). 310 Collectively, the crosscheck of the mathematical model with the RNA-seq data revealed 311 predominantly the agreement of decreased enzyme activities with reduced enzyme gene 312 expression in TCAIM KI CD8+ T cells. Moreover, TCAIM KI CD8+ T cells seem to fail to 313 increase their ΔΨm as well as their HIF1α protein expression upon activation further 314 confirming our assumption of impaired metabolic activity and thus, effector differentiation due 315 to TCAIM KI. 316 317 TCAIM inhibits cholesterol biosynthesis in activated CD8+ T cells 318 As we observed a regulatory role of TCAIM for metabolic adaptations upon T cell activation, 319 we investigated whether there was a mechanistic explanation evident from our RNA-seq 320 data. In a non-supervised hierarchical clustering analysis naïve CD8+ T cells clearly 321 separated from all activated T cells. TCAIM KI or KO had no further effect on the 322 subclustering of naïve CD8+ T cells. Interestingly, the expression profile of 48 h activated 323 TCAIM KI CD8+ T cells remained more similar to that of 24 h activated TCAIM KI and wt 324 CD8+ T cells, whereas 48 h activated TCAIM KO and wt CD8+ T cells acquired a distinct 325 expression profile (Fig. 5a). 326 To understand the functional differences associated with the altered expression profile in 327 activated TCAIM KI CD8+ T cells, we performed a gene ontology (GO) enrichment analysis in 328 which genes were differentially expressed between 48 h activated TCAIM KI and wt CD8+ 329 T cells (Fig. 5b). The majority of annotations found among the GO terms describing biological 330 processes were related to immunologic processes including cytokine production, responses 331 to stimuli or cell differentiation. Genes that repeatedly contributed to that annotations 332 comprised e.g. Klf2, Il-1r2, Traf3ip1, Ifng-γ, Klrk1, IL-23a, Cd80 or Il-12rb (Extended Data 333 Tab. 1) with the latter five as markers for T cell activation and effector function being 334 downregulated, while the others as markers for T cell quiescence and inhibition being 335 upregulated in TCAIM KI CD8+ T cells. This is again consistent with the impaired T cell 336 activation and effector function due to TCAIM KI. 337 Remarkably, apart from the regulation of genes controlling immunologic processes we 338 detected a strong downregulation of genes involved in cholesterol biosynthetic processes. 339 Genes within this set were essential enzymes of the mevalonate pathway (Fig. 5c) and the 340 cholesterol biosynthesis which branches off the mevalonate pathway (Fig. 5d). Indeed, 341 TCAIM KI CD8+ T cells either failed to upregulate gene expression at all (e.g Mvd or Lss) or 342 to further increase gene expression after 24 h of activation (Pmvk, Fdps, Fdft1, Sqle, Cyp51, 343 Msmo1 and Nsdhl; Fig. 5c,d). This was also true for most of the remaining enzymes of the 344 mevalonate and cholesterol biosynthesis pathways although their changes were not 345 significant (Extended Data Fig. 4a,b). 9

bioRxiv preprint doi: https://doi.org/10.1101/2021.04.20.440500; this version posted April 21, 2021. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY-NC-ND 4.0 International license.

346 Taken together, the RNA-seq analysis further confirmed an impaired T cell activation and 347 acquisition of effector functions and additionally, revealed a deficiency in expression of 348 cholesterol biosynthesis genes due to TCAIM KI. 349 350 Cholesterol supplementation restores activation and proliferation of TCAIM KI CD8+ 351 T cells 352 As TCAIM KI CD8+ T cells showed a reduced cholesterol biosynthesis based on the 353 RNA-seq data, we tested whether replenishment of free cholesterol could restore CD8+ T cell 354 activation and especially effector differentiation. To investigate this, naïve CD8+ T cells with 355 homozygous, heterozygous TCAIM KI or wt TCAIM gene expression (hocre, hecre or wtcre, 356 respectively) were cultured in presence of increasing amounts of cholesterol for up to 72 h. 357 We then performed flow cytometry analysis and measured CD62L, CD44 and HIF1α protein 358 expression as well as CD8+ T cell proliferation. 359 CD8+ T cell activation and differentiation clearly depended on the level of TCAIM 360 overexpression. The majority of untreated hocre TCAIM KI CD8+ T cells again maintained a 361 naïve phenotype expressing CD62L only, whereas untreated hecre TCAIM KI CD8+ T cells 362 showed an increase in central memory and effector/ effector memory frequencies. Those, 363 however, still were below that of untreated wtcre CD8+ T cells. Remarkably, this changed 364 upon addition of cholesterol in a dose-dependent manner. We detected a significant increase 365 in central memory cell frequencies upon treatment with the highest cholesterol dose for hocre 366 TCAIM KI CD8+ T cells, which were then comparable to the levels of untreated wtcre CD8+ 367 T cells. A similar tendency although not significant could also be observed for the effector/ 368 effector memory cell pool of hocre as well as hecre TCAIM KI CD8+ T cells. Wtcre CD8+ 369 T cells, however, did not respond to the cholesterol treatment and showed no further 370 activation (Fig. 6a,b). The increased activation upon cholesterol treatment was accompanied 371 by a likewise increased HIF1α protein expression and proliferation rate of both hocre and 372 hecre TCAIM KI but not wt CD8+ T cells (Fig. 6c-e). 373 In summary, the addition of exogenous cholesterol was sufficient to restore the activation, 374 proliferation and partially differentiation deficiency of TCAIM KI CD8+ T cells. Thus, we 375 conclude that TCAIM mediated defects in CD8+ T cell activation and differentiation were due 376 to a decrease in cholesterol biosynthesis enzyme expression. 377

378 TCAIM controls activation-induced MERC formation 379 Mitochondrial fission accompanied cristae remodeling was proposed as prerequisite for 380 activation-induced metabolic adaptations and finally acquisition of effector function13. As 381 TCAIM KI CD8+ T cells showed an altered metabolic program and impaired effector 382 differentiation, we next investigated whether TCAIM regulates mitochondrial dynamics and 383 cristae remodeling. Thus, we analyzed electron microscopy images of naïve (0h) and 384 24 - 48 h polyclonally activated TCAIM KI, KO and wt CD8+ T cells (Fig. 7a and Extended 385 Data Fig. 5a). 386 The data revealed an increase in the average number of mitochondria per cell upon 387 activation of wt but also TCAIM KI as well as TCAIM KO CD8+ T cells. This is in line with 388 previous findings on human and murine CD8+ T cells, which showed that the number of 389 mitochondria per cell increased upon activation24, 25. However, while the increase in number 390 of mitochondria was fast but only transient for wt and TCAIM KO CD8+ T cells, TCAIM KI 391 CD8+ T cells showed a delayed increase in mitochondrial number (Extended Data Fig. 5b). 392 Against expectations, no structural differences in mitochondrial shape towards a more 393 elongated morphology pointing towards a less anabolic phenotype could be observed for 10

bioRxiv preprint doi: https://doi.org/10.1101/2021.04.20.440500; this version posted April 21, 2021. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY-NC-ND 4.0 International license.

394 TCAIM KI CD8+ T cells. Also vice versa, TCAIM KO did not cause an increased fission rate in 395 TCAIM KO CD8+ T cells (Extended Data Fig. 5c). However, in line with the increase in 396 average mitochondrial number, we observed an activation-dependent transient increase in 397 cristae formation for all TCAIM KI, KO and wt CD8+ T cells (Extended Data Fig. 5d). 398 Interestingly, we detected disparities of MERCs between activated TCAIM KI, KO and wt 399 CD8+ T cells (Fig. 7a). On average, 60 % of the mitochondria from a single naïve CD8+ T cell 400 were in contact with the ER. This increased dramatically upon activation with up to 90 % of 401 the mitochondria from wt and TCAIM KO CD8+ T cells engaging contact with the ER 24 h 402 after stimulation (Fig. 7b). In contrast, mitochondria from TCAIM KI CD8+ T cells displayed a 403 delayed MERC response with 70 % and 90 % of the mitochondria forming contacts at 24 and 404 48 h, respectively. Whereas, MERCs of wt and most strikingly TCAIM KO CD8+ T cells 405 began to dissolve 48 h after stimulation as only 85 % and 70 % of the mitochondria, 406 respectively, had MERCs left (Fig. 7b). Also, the number as well as the length of single ER 407 tubule being in contact with a increased upon activation (Fig. 7c,d). 408 Importantly, TCAIM KO CD8+ T cells had significantly less ER tubule contacts per 409 mitochondrion (Fig. 7c) while the ER tubule contacts formed within TCAIM KI CD8+ T cells 410 had a highly increased length and shape with some mitochondria being nearly completely 411 covered by ER tubules (Fig. 7d) after 48 h of stimulation. 412 Collectively, activated TCAIM KI CD8+ T cells showed a delayed but strong increase in 413 number of mitochondria and delayed formation of very long MERCs. No clear evidence for 414 promotion of mitochondrial fusion was found due to TCAIM KI. Thus, the altered anabolic 415 adaptation of activated TCAIM KI CD8+ T cells may not be caused by changed mitochondrial 416 dynamics in general but by delayed and altered MERC formation. 417 418 TCAIM interacts with proteins being enriched at MERC sites 419 The interaction between mitochondria and ER has been described to regulate mitochondrial 420 metabolism and bioenergetics, mitochondrial fission, lipid metabolism, cellular Ca2+ signaling, 421 autophagosome formation and apoptosis26, 27. To mediate these processes, mitochondria 422 and ER membranes need to physically tether. This is accomplished by specific protein 423 interactions, which are not yet fully identified for mammalian and especially immune cells. 424 Protein pairs that were previously characterized to actively mediate mitochondria-ER 425 tethering include mitochondrial MFN1 and MFN2 interacting with ER bound MFN2, VDAC2 426 with IP3R and Gpr75 or RMD3 with VAPB, respectively 27. Thus, to get a more detailed 427 information on how TCAIM interferes with MERC formation we identified its interaction 428 partners. Hence, we performed a anti-GFP co-immunoprecipitation with cell lysates from 429 HEK293T cells transfected with either a plasmid encoding for a TCAIM-eGFP fusion protein 430 or an eGFP protein with mitochondrial targeting sequence as a control. Subsequent mass 431 spectrometry of the retained co-immunoprecipitates revealed only 14 TCAIM specific 432 interaction partners (Fig. 8a). These include TCAIM itself suggesting formation of homomers 433 to exert TCAIM specific functions. The other identified interaction partners were mostly 434 proteins known to be located to mitochondria. More specifically, six of the TCAIM interaction 435 partner, including TOM40, SAM50, VDAC2, MIC60, DJC11 and STML2, belong to the 436 mitochondrial contact site and cristae organizing system (MICOS) protein complex and 437 MICOS-associated proteins. MICOS and MICOS-associated proteins have been shown to 438 regulate cristae assembly, but also, to participate in MERC formation, where they mediate 439 the segregation of inner mitochondrial membrane components for appropriate cristae, protein 440 and mtDNA distribution28. Notably, the previously characterized mitochondria-ER tethering 441 proteins VDAC2 and RMD3 were among the identified TCAIM interaction partners. 442 Gene expression of both proteins (Rmdn3 encodes for RMD3) was upregulated upon 443 activation in all TCAIM KI, KO and wt CD8+ T cells, whereas TCAIM gene expression was 11

bioRxiv preprint doi: https://doi.org/10.1101/2021.04.20.440500; this version posted April 21, 2021. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY-NC-ND 4.0 International license.

444 only downregulated in TCAIM KO and wt but not TCAIM KI CD8+ T cells (Fig. 8b). Despite 445 VDAC2 and RMD3, also gene expression of SAM50 and MIC60, core-subunits of the MICOS 446 protein complex, was upregulated in TCAIM KI, KO and wt CD8+ T cells (Fig. 8c). 447 Finally, bimolecular fluorescence complementation (BiFC) assay with subsequent flow 448 cytometric analysis was used to confirm interaction of TCAIM with the MICOS subunit 449 TOM40. The assay is based on the reconstitution of a fluorescent protein. HEK293T cells 450 were transfected with two plasmids each encoding for the indicated protein of interest fused 451 to either the N- or C-terminal part of eGFP. In case of protein interaction, the complementary 452 eGFP fragments refold and start to fluoresce. Transfection of TCAIM fused to C-terminal part 453 of eGFP alone was used as control for false positive signals like auto fluorescence of 454 HEK293T cells while transfection of plasmids encoding for TOM40 and MIC60 which are 455 known to interact with each other29 within the MICOS complex served as positive control. 456 Indeed, eGFP positive cells were observed when complementary eGFP fragments coupled 457 to TCAIM and TOM40 as well as TOM40 and MIC60 were transfected (Fig. 8d). 458 In conclusion, our results revealed that TCAIM interacts with proteins promoting MERC 459 formation and function. Moreover, our results show that T cell activation enhances 460 transcription of MERC forming proteins, whereas, down-regulation of TCAIM expression 461 might circumvent inhibition of MERC formation. 462 463

12

bioRxiv preprint doi: https://doi.org/10.1101/2021.04.20.440500; this version posted April 21, 2021. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY-NC-ND 4.0 International license.

464 Discussion 465 T cells need to adapt their cellular metabolism for full effector cell differentiation upon 466 activation and it has been reported that this relies on alterations in mitochondrial physiology 467 including cristae resolution, fission and an increase in ΔΨm and mROS production11, 13. 468 However, the molecular signals driving mitochondrial changes and the communication 469 between mitochondria and the ER thereby mediating changes in cellular metabolism remain 470 less clear. We recently reported that the mitochondrial protein TCAIM is downregulated upon 471 T cell activation and inhibits effector cell differentiation of conventional CD4+ T cells15. Our 472 previous investigations also indicated that TCAIM interferes with activation-induced changes 473 in mitochondrial morphology and function15. Thus, TCAIM seems to function as an important 474 link between activation-induced mitochondrial alterations and changes in cellular metabolism 475 driving full effector cell differentiation. 476 Therefore, we here utilized conditional TCAIM KI and KO mice as a model to investigate 477 which molecular processes in the mitochondria control effector T cell differentiation: First, we 478 could show that TCAIM KI causes also in CD8+ T cells comparable effects of impaired 479 proliferation, effector/ effector memory differentiation and cytokine production as we 480 previously reported for CD4+ T cells15. Second, our data on TCAIM KO CD8+ T cells point 481 towards an advantage for effector/ effector memory differentiation in vivo. This further 482 supports the regulatory role of TCAIM for activation and differentiation of T cells in general. 483 Third, metabolomics and transcriptional data revealed that TCAIM KI inhibits 484 activation-induced enforcement of TCA cycle activity and glutaminolysis, but more 485 importantly, that TCAIM KI blocks upregulation of the mevalonate pathway and the 486 cholesterol biosynthesis. Indeed, addition of cholesterol rescued T cell activation and 487 proliferation of TCAIM KI CD8+ T cells. Finally, at the molecular level, we could show that 488 TCAIM delays activation-induced interaction between mitochondria and the ER probably by 489 binding to proteins known to promote MERC formation such as RMD3 and VDAC. 490 For a long time, many studies on activation-induced metabolic adaptations in T cells were 491 primarily focusing on the glycolysis and TCA cycle axis as the central carbon processing 492 pathways. However, enforcement of lipid synthesis following activation has been shown to be 493 likewise important for T cell proliferation and acquisition of effector functions. Chen et al.30 494 already found in the 1970th, a strong upregulation of cholesterol biosynthesis in lymphocytes 495 within the first 24 h upon activation and showed the importance for induction of DNA 496 replication and cell proliferation. Later, in the 1990th, Chakrabarti et al.31 showed, that T cells 497 were unable to proceed in cell cycle progression upon inhibition of cholesterol biosynthesis 498 and instead remained in G1 phase. In addition to promoting T cell proliferation, Kidani et al.32 499 reported on the importance of cholesterol stores and biosynthesis for virus specific effector 500 cell development, whereas Yang et al.33 demonstrated the need of increased membrane 501 cholesterol during T cell activation for enhancement of TCR signaling and immunologic 502 synapse formation. 503 Our study not only supports the importance of cholesterol for CD8+ T cell activation and 504 proliferation, but also revealed a crucial role of activation-induced MERC formation which 505 was inhibited by TCAIM KI. Notably, cholesterol biosynthesis enzymes are located in MERC 506 structures34 further highlighting the role of mitochondria-ER communication for cholesterol 507 biosynthesis. We could show that TCAIM binds to MERC promoting proteins such as VDAC 508 and RMD3, which are crucial for MERC formation and function including lipid biosynthesis35, 509 36, 37, 38.

510 Moreover, we showed that TCAIM interferes with increased HIF1α protein expression upon 511 T cell activation, which could be restored upon addition of exogenous cholesterol. 512 Cholesterol treatment of hepatocytes has been shown to induce HIF1α protein accumulation 513 mediated by inhibition of HIF1α protein degradation39. Furthermore, HIF1α protein expression 13

bioRxiv preprint doi: https://doi.org/10.1101/2021.04.20.440500; this version posted April 21, 2021. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY-NC-ND 4.0 International license.

514 as well as induction of CD44 is promoted by mammalian target of rapamycin complex 1 515 (mTORC1)40, 41. MTORC1 also promotes SREBP2 activation42, 43 and protein expression32 516 which is crucially controlling transcription of mevalonate pathway and cholesterol 517 biosynthesis enzymes44. Indeed, we found reduced cholesterol biosynthesis enzyme gene 518 expression as well as HIF1α and CD44 protein expression in activated TCAIM KI CD8+ 519 T cells suggesting a deficiency in mTORC1 signaling through TCAIM mediated mechanisms. 520 Although not explicitly shown for T cells, mTORC1 has been found to be located to 521 mitochondria thereby regulating mitochondrial functions45. Reversely, Head et al.46 revealed 522 that mTOR activation is dependent on VDAC1 function. VDACs are considered as regulatory 523 elements of metabolic crosstalk between mitochondria and the cytosol47. There are three 524 isoforms: VDAC1, VDAC2 and VDAC3. Whereas VDAC1 and VDAC2 are assumed to share 525 similar functions in regulating the exchange of Ca2+, ATP/ ADP and other metabolites, 526 VDAC3 seems to be functionally distinct. Notably, deletion of VDAC2, the isoform TCAIM is 527 interacting with, is the only embryonically lethal one and overexpression of VDAC1 and 528 VDAC3 fail to compensate for VDAC2 KO47 highlighting its non-redundant role. Furthermore, 529 it has been described that binding of mTORC1 to VDAC1 and Bcl-xl enhances ΔΨm 45, 48. 530 Interestingly, TCAIM KI CD8+ T cells show a reduced activation-induced upregulation of 531 ΔΨm which might be caused by TCAIM-VDAC2 interaction. 532 Our investigations not only revealed delayed MERC formation in TCAIM KI T cells but also 533 an increased length of these late forming subcellular structures. Interestingly, long MERCs 534 have been observed in situations when the mTORC1 nutrient-sensing pathway disengages49. 535 Thus, indeed TCAIM interaction with VDAC might interfere with mTORC1 activation at the 536 mitochondria resulting in nutrient-drop-mediated delayed enlargement of MERCs. 537 Furthermore, MERCs and especially the local activity of mTOR and VDAC at MERCs have 538 been previously reported to function as immunometabolic hubs linking TCR signals to 539 enforcement of glycolysis50. Subcellular accumulation of mTORC2-AKT-GSK3 at MERCs 540 enabled fast recruitment of hexokinase I to VDAC and thus, induction of metabolic flux into 541 mitochondria. However, this was exclusively shown for rapid memory CD8+ T cell responses 542 and limited to glycolytic flux. With our study, we additionally demonstrate an essential role of 543 MERCs and TCAIM for cholesterol biosynthesis and effector cell differentiation during 544 primary T cell activation. 545

14

bioRxiv preprint doi: https://doi.org/10.1101/2021.04.20.440500; this version posted April 21, 2021. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY-NC-ND 4.0 International license.

546 Methods 547 Mice. TCAIM KO mice were purchased from MRC Harwell Institute, C57BL/6N mice from 548 Charles River Laboratories, C57BL/6-Tg(CAG-Flpe)2Arte mice from TaconicArtemis GmbH 549 and Cd4Cre-Car mice were provided by M. Schmitt-Supprian. TCAIM KI Cd4Cre mice were 550 previously described15. TCAIM KO mice were crossed first with C57BL/6-Tg(CAG-Flpe)2Arte 551 and subsequent with Cd4Cre-Car mice to obtain TCAIM KO Cd4Cre mice bearing a 552 conditional T cell specific deletion of exon four of the Tcaim gene resulting in a loss of protein 553 function. TCAIM KI Cd4Cre and TCAIM KO Cd4Cre mice strains were on a C57BL/6 554 background. All mice were bred and/ or housed under specific pathogen-free conditions in 555 the animal facility of the Charité. For experiments, gender-matched male and female 556 littermates or mice from parallel breeding (max. 1 - 7 weeks age divergent) aged between 557 2 - 6 months were used. Studies were performed in accordance with the guidelines of the 558 Federation of European Laboratory Animal Science Association (FELASA) and approved by 559 the local authority [Landesamt für Gesundheit und Soziales Berlin (LAGeSo)]. 560 561 Preparation of single cell suspensions. Spleen and lymph nodes were first forced through 562 100 µm cell strainers (Falcon/ Corning) with 4 °C cold phosphate buffered saline (PBS, 563 Gibco) and centrifuged at 4 C and 300 g. Then erythrocytes of spleen cell suspensions were 564 lysed in hypotonic PBS diluted in a 1:3 ratio with sterile deionized H2O for 12 seconds. Lysis 565 was stopped by addition of PBS supplemented with 2 % (vol/ vol) fetal calf serum (FCS, 566 Biochrom). Lastly, single cell suspensions of spleen and lymph nodes were filtered through 567 40 µm cell strainers (Falcon/ Corning) with 4 °C PBS containing 2 % (vol/ vol) FCS. 568 For isolation of colon cells, the entire colon was isolated, rinsed with ice cold Hanks' 569 Balanced Salt Solution (HBSS) without calcium and magnesium (Gibco) supplemented with 570 2 % (vol/ vol) FCS and and 10 mM 4-(2-Hydroxyethyl)piperazine-1-ethanesulfonic acid 571 (HEPES, Merck) and cut into 1 cm pieces. Colon pieces were placed in HBSS containing 572 2 % (vol/ vol) FCS, 10 mM HEPES and 0,154 mg/ ml 1,4-Dithioerythritol (DTE, Sigma) and 573 incubated in a shaker incubator (200 rpm, 37 °C) for 15 minutes. This was repeated twice. 574 Then colon pieces were placed in HBSS containing 2 % (vol/ vol) FCS, 10 mM HEPES and 575 5 mM Ethylenediaminetetraacetic acid (EDTA, Sigma-Aldrich) and incubated in in a shaker 576 incubator (500 rpm, 37 °C) for 15 minutes. This was repeated twice. Residual EDTA was 577 removed by rinsing colon pieces in VLE RPMI 1640 with glutamine (Biochrom AG). Then 578 colon pieces were digested in RPMI 1640 containing 10 % (vol/ vol) FCS, 0,1 mg/ ml 579 Liberase TL Research Grade (Roche) and 0,1 mg/ ml DNAse I (Roche) for 20 minutes in a 580 shaker incubator (200 rpm, 37 °C). Following incubation, remaining tissue pieces were first 581 forced through 100 µm cell strainers with HBSS containing 10 mM HEPES and then filtered 582 through 40 µm cell strainers with VLE RPMI 1640 with glutamine containing 10 % (vol/ vol) 583 FCS, 100 U/ mL penicillin and 100 mg/ mL streptomycin (all Biochrom AG). 584 585 Cell culture and cell lines. Naive CD8+ T cells were isolated from single cell suspensions of 586 spleen and lymph nodes using a Naive CD8+ T Cell Isolation Kit, mouse (Miltenyi) or 587 EasySep Mouse Naïve CD8+ T Cell Isolation Kit (StemCell) and cultured in VLE RPMI 1640 588 with glutamine supplemented with 10 % (vol/ vol) FCS, 100 U/ mL penicillin, 100 mg/ mL 589 streptomycin (all Biochrom AG) and 50 µM -mercaptoethanol (Sigma-Aldrich) at 37 °C and 5 590 5 % CO2. If not stated otherwise, 1 x 10 cells were polyclonally stimulated with 2 µg/ ml 591 plate-bound αCD3 (145-2C11, eBioscience) and 1 µg/ ml soluble αCD28 (37.51, 592 eBioscience) for the indicated time. To quantify CD8+ T cell proliferation, cells were stained 593 with Cell Proliferation Dye eFluor™ 450 (eBioscience) according to manufacturer’s protocol 594 prior culture. If indicated, cyclosporin A (Merck), DMSO (Sigma-Aldrich) or cholesterol– 15

bioRxiv preprint doi: https://doi.org/10.1101/2021.04.20.440500; this version posted April 21, 2021. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY-NC-ND 4.0 International license.

595 methyl- -cyclodextrin (Sigma-Aldrich) was added to the culture in increasing concentrations 596 (0.01 – 0.5 µM, 10 µM, or 2 – 40 µM, respectively). For some experiments, naïve CD8+ 597 T cells had been co-cultured with autologous B cells isolated from spleen cell suspensions 598 using an EasySep Mouse B Cell Isolation Kit (StemCell) in a 1:1 ratio. 599 HEK293T cell line (ACC 305) was purchased from the DSMZ-German Collection of 600 Microorganisms and Cell Cultures GmbH and maintained in DMEM (very low endotoxin, 601 Biochrom AG and PAN Biotech) supplemented with 10 % (vol/ vol) FCS, 100 U/ mL 602 penicillin, 100 mg/ mL streptomycin at 37 °C and 5 % CO2. 603 604 Non-targeted metabolite analysis. Sample preparation and GC/APCI/MS (gas 605 chromatographic atmospheric pressure chemical ionization mass spectrometry) analysis was 606 performed as described previously51. Briefly, metabolites of pellets of 7 x 105 60 h 607 polyclonally activated CD8+ T cells and 10 µl of supernatants from 0, 12, 20, 36, 44 and 60 h 608 CD8+ T cell cultures were extracted by incubation with 90 µl ice cold methanol for 30 min 609 under shaking at 4 °C. Subsequently, 50 µl H2O was added, samples were vortexed and 610 centrifuged at 21,000 g for 5 min at room temperature. Supernatants were vacuum dried and 611 stored at -80 °C until measurement. For measurement, the dried mixtures were derivatized 612 on-line using a PAL RTC autosampler (CTC Analytics). After addition of 10 μl methoxyamine 613 (20 mg/ ml in pyridine; Sigma), vials were agitated for 90 min at 34 °C and 750 rpm. 90 μl 614 N-methyl-N-(trimethylsilyl)-trifluoroacetamide (MSTFA; Macherey-Nagel) containing 615 0.2 μg/ ml each of C4-C24 fatty acid methyl esters (Sigma) as retention index markers were 616 added, followed by agitating vials for 30 min. Before injection, samples were allowed to rest 617 for 2 h to complete derivatization reactions. GC-APCI-MS analysis was carried out on an 618 Agilent 7890 B gas chromatograph (Agilent) coupled to an Impact II quadrupole time-of-flight 619 (Q-TOF) mass spectrometer (Bruker) via a GC-APCI II source (Bruker). 1 μl sample was 620 injected into a split/ splitless inlet, operated at 230 °C in split mode (1:10). Chromatographic 621 separation was carried out on a 30 m x 0.25 mm x 0.25 μm HP5-MS UltraInert capillary 622 column (Agilent) connected to a 0.5 m x 0.25 mm RxiGuard (Restek) column as transfer 623 capillary. Helium was used as a carrier gas at 1 ml/ min. The temperature gradient was as 624 follows: 80 °C held for 2 min, 15 °C/ min to 320 °C, 320 °C held for 7 min. Transfer capillary 625 and capillary head of the GC-APCI II source were heated to 290 °C. The GC-APCI source 626 was operated in positive ion mode, using a capillary voltage of -2500 V and an end plate 627 offset of -250 V. A corona current of 3000 nA and a nebulizer pressure of 2.0 bar was used. 628 The dry gas was nitrogen at a flow rate of 2.0 l/ min and a temperature of 230 °C. Full-scan 629 line spectra were recorded in the scan range of m/ z 80 - 1000 at an acquisition rate of 10 s-1. 630 Q-TOF ion transfer parameters were as follows: funnel 1 RF 300.0 Vpp, funnel 2 RF 631 300.0 Vpp, isCID energy 0.0 eV, hexapole RF 100.0 Vpp, quadrupole ion energy 4.0 eV, low 632 mass 80.0 m/ z, collision energy 8.0 eV, collision RF 1000.0 Vpp, transfer time 75 μs, 633 pre-pulse storage 8.0 μs. External mass calibration was performed before each sequence by 634 direct infusion of 180 μl/ h ES low concentration tune mix (Agilent) according to 635 manufacturer’s instruction. Subsequently, metabolites were identified in comparison to the 636 Golm Metabolome Database52 or putatively annotated as described previously53. Following 637 identification, metabolite peak intensities were log10 transformed and normalized for 638 experimental replication and run order effects using an ANOVA based method54. 639 Log-transformation improves normality of the data distribution. Run order effects are 640 systematic effects on measured metabolite intensities due to the measurement time point. 641 Experimental replications describe the samples being generated in three independent 642 laboratory trials. ANOVA based normalization removes effects related to these factors prior 643 to further statistical tests. 644

645 Flow cytometry. Dead cells were stained using either Zombie UV™ Fixable Viability Kit 646 (Biolegend, 1:1000), Fixable Viability Dye eFluor™ 506 (eBioscience, 1:100) or 16

bioRxiv preprint doi: https://doi.org/10.1101/2021.04.20.440500; this version posted April 21, 2021. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY-NC-ND 4.0 International license.

647 LIVE/DEAD™ Fixable Far Red Dead Cell Stain Kit (Invitrogen, 1:5000) according to 648 manufacturer’s protocol. Concomitantly, Fc receptors were blocked with Purified Rat Anti- 649 Mouse CD16/CD32 Mouse BD Fc Block™ (2.4G2, BD Pharmingen, 1:1000). Cell surface 650 antigens were stained in PBS containing 2 % (vol/ vol) FCS and 1 mg/ ml sodium azide 651 (Serva) at 4 °C in the dark. Cell fixation and staining of intracellular antigens was done with 652 Foxp3 Staining Buffer Set (Miltenyi) according to manufacturer’s protocol. For cytokine 653 detection, cells were treated prior staining with 10 ng/ mL phorbol 12-myristate 13-acetate 654 (Sigma-Aldrich) and 1 µg/ mL ionomycin (Biotrend) for 4 h at 37 °C. For the last 2 h 2 µg/ mL 655 Brefeldin A (Sigma-Aldrich) was added. Monoclonal antibodies against the following antigens 656 were used: CD3 (17A2, Biolegend, CD8a (53-6.7, Biolegend or BD Pharmingen), CD25 657 (PC61, Biolegend), CD44 (IM7, Biolegend), CD45R (RA3-6B2, Biolegend or Miltenyi), 658 CD62L (MEL-14, Biolegend), CD69 (H1.2F3, Biolegend), GZMB (NGZB, eBioscience), Hif1α 659 (Mgc3, eBioscience), IFN- (XMG1.2, Biolegend), Ki67 (SolA15, eBioscience) ), TCR 660 (H57-597, Biolegend). For measurement of ΔΨm or ROS production, cells were stained with 661 0.1 µM MitoStatus TMRE (BD Pharmingen) for 20 min or 5 µM MitoSOX™ Red mitochondrial 662 superoxide indicator (Invitrogen) for 10 min at 37 °C according to manufacturer’s protocol, 663 respectively, and subsequently, stained for dead cells and surface antigens simultaneously in 664 PBS or HBSS with Ca2+/ Mg2+ (Gibco) for 10 min at 4 °C. Data were acquired using an 665 LSRFortessa with BD FACSDiva software v8.0.2 (BD Biosciences) or a CytoFLEX LX with 666 CytExpert software v2.3.1.22 (Beckman Coulter) and analyzed with FlowJo v10 software 667 (FlowJo, LLC). 668 669 Immunohistology. Colon samples were fixed in 36.5 % formaldehyde and embedded in 670 paraffin. Sections of 1 – 2 µm thickness were cut, deparaffinated and subjected to a 671 heat-induced epitope retrieval step. Endogenous peroxidase was blocked by hydrogen 672 peroxide prior to incubation with αCD3 (polyclonal rabbit, Agilent #IR50361-2) followed by the 673 EnVision+ System-HRP Labelled Polymer Anti-Rabbit (Agilent). For visualization, OPAL-570 674 diluted in amplifier diluent (Akoya Biosciences) was used. Proteins were then inactivated and 675 sections incubated with αCD4 (4SM95, eBioscience) followed by incubation with rabbit 676 anti-rat secondary antibody (Dianova), then with the EnVision+ System-HRP Labelled 677 Polymer Anti-Rabbit and finally with OPAL-670 diluted in amplifier diluent (Akoya 678 Biosciences). The procedure was repeated for αCD8 (4SM15, eBioscience) staining using 679 OPAL-520 (Akoya Biosciences) for visualization. Nuclei were stained using 4′,6-diamidine-2′- 680 phenylindole dihydrochloride (DAPI, Sigma). Slides were coverslipped in Fluoromount G 681 (Southern Biotech). Multispectral images were acquired using a Vectra® 3 imaging system 682 employing the Vectra3 and Phenochart software (all Akoya Biosciences). Images are 683 displayed in pseudocolours. Cells were quantified in a blinded manner using the inForm 684 Software (Akoya Biosciences). 685 686 Quantitative real time PCR. RNA of polyclonally activated CD8+ T cells from C57BL6/N 687 mice was purified using NucleoSpin RNA kit (Macherey-Nagel) according to manufacturer’s 688 protocol. RNA was then incubated at 65 °C for 5 min. Subsequently, cDNA was synthesized 689 using QuantiTect Reverse Transcription kit (Qiagen). Gene expression was quantified using 690 TaqMan gene expression assays (Thermo Fisher) for Tcaim and Ubc or custom designed 691 primer and probes for Il-2 and Ifn-γ (Eurofins Genomics) and TaqMan Universal PCR Master 692 Mix (Thermo Fisher) on a 7500 Real Time PCR System (Applied Biosystems). Sequences 693 for murine Il-2: forward primer: TCGCCAGTCAAGAGCTTCAGACAAGCA, reverse primer: 694 CATGCCGCAGAGGTCCAA, probe: CAATTCTGTGGCCTGCTTGGGCAA. Sequences for 695 murine Ifn-γ: forward primer: AGCAACAGCAAGGCGAAAAA, reverse primer: 17

bioRxiv preprint doi: https://doi.org/10.1101/2021.04.20.440500; this version posted April 21, 2021. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY-NC-ND 4.0 International license.

696 AGCTCATTGAATGCTTGGCG, probe: ATTGCCAAGTTTGAGGTCAACAACCCACA. 697 Probes were coupled to FAM and TAMRA. The thermal cycling program was set as followed: 698 2 min at 50 °C, 10 min at 95 °C and 40 cycles of 15 s at 95 °C and 1 min at 60 °C. The 699 reactions were performed as duplicates. Data were analyzed with 7500 Systems SDS 700 Software v2.3 and relative gene expression was calculated as 2-ΔCT using Ubc as reference 701 gene. 702 703 RNA-seq. RNA of naïve and polyclonally activated CD8+ T cells was purified using 704 NucleoSpin RNA kit (Macherey-Nagel) according to manufacturer’s protocol and submitted to 705 the Scientific Genomics Platform (Max Delbrück Center for Molecular Medicine) for 706 sequencing. Briefly, mRNA-Seq libraries were prepared from 100 ng of input total RNA 707 samples using TruSeq stranded mRNA kit (Illumina). Barcoded and pooled libraries were 708 sequenced on 5.5 lanes of Illumina HiSeq 4000 platform (HCS v3.3.76, RTA v2.7.6) in a 709 paired-end 2 x 75 nt sequencing mode. Fastq files generation and demultiplexing were 710 performed using bcl2fastq v2.20.0. 711 712 RNA-seq raw data processing. The quality of fastq files was controlled by FastQC v0.11.8 713 (Bioinformatics Group at the Babraham Institute). Residual adapter sequences, ambiguous 714 bases, low quality reads and reads with a length below 25 were removed by 715 AdapterRemoval v2.2.355. The remaining reads were aligned to the GRCm38 Ensembl 716 version of the mm10 mouse genome using Tophat v2.1.156 and Bowtie2 v2.2.557. Counts per 717 gene were summarized by the featureCount algorithm of the Rsubread package v2.4.258 in R 718 v4.0.2 (R Core Team (2020). R: A language and environment for statistical computing. R 719 Foundation for Statistical Computing, Vienna, Austria. URL https://www.R-project.org/.). 720 721 RNA-seq data analysis. Raw count data of protein-coding genes were normalized and 722 either log2-transformed after adding a pseudocount to each value or variance-stabilized 723 transformed using the DESeq2 package v1.30.059 in R. 1000 genes with the highest 724 variances across all samples were selected and z-scores of the normalized expression 725 values of these genes were subjected to hierarchical clustering. The heatmap representation 726 of the clustering result was done with the in-built heatmap function in R. Z-scores of 727 expression values of further genes preselected by functional aspects were shown in a 728 heatmap with predefined order created with the pheatmap package v1.0.12 (Raivo Kolde 729 (2019). pheatmap: Pretty Heatmaps. R package version 1.0.12. https://CRAN.R- 730 project.org/package=pheatmap). 731 Differential expressed genes between two groups were determined by fitting models of 732 negative binomial distributions to the count data. Raw p-values were adjusted for multiple 733 testing using false discovery rate. Genes with significant differential expression were 734 selected by an adjusted p-value below 0.05 and a minimal absolute log2-fold change of one. 735 Enrichment of functional aspects within the sets of differential expressed genes were 736 analyzed after mapping the genes to functions by DAVID using the R-packages 737 RDAVIDWebService v1.26.060 and clusterProfiler v3.16.161. Selected results were plotted 738 using GOCircle - function of the GOplot package v1.0.262. 739 740 Mathematical metabolite flux modelling. We modeled glycolysis, serine biosynthesis, TCA 741 cycle, and glutaminolysis pathways as they represent the main carbon flux routes for 742 biomass precursors and energy production, therefore, reflecting the major changes in 18

bioRxiv preprint doi: https://doi.org/10.1101/2021.04.20.440500; this version posted April 21, 2021. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY-NC-ND 4.0 International license.

743 metabolic activity. The overall network was constructed to have at maximum two 744 unmeasured intermediate metabolites except in reactions from glycolysis or TCA cycle. All 745 carbon molecules taken up were either converted into biomass building blocks or secreted. 746 To estimate the biomass of murine CD8+ T cells, we used an adapted version of Thiele's 747 human lymph node germinal center biomass reaction as the closest available genome scale 748 metabolic reconstruction63. Growth based dilution was negligible in this model as the fluxes 749 were orders of magnitude higher than the dilution fluxes. The relative measurements from 750 the non-targeted metabolite analysis of TCAIM KI and wt CD8+ T cells and their culture 751 supernatants were converted to absolute values using calibration data from Lisec et al.64 752 combined with log-log-regression. The kinetics of the model were either reversible or 753 irreversible mass action kinetics. The kinetic rates of TCAIM KI CD8+ T cells were considered 754 as deviation from the kinetic rates of wt CD8+ T cells and, therefore, expanded with 755 regulation parameters. All co-factors were held constant at a value of 1 mM as they were not 756 measured in the experiments. We assumed steady state for the entire model. The objective 757 of the optimization was the minimization of the sum of the absolute values of the logarithm of 758 regulation parameters, representing a minimal metabolic change between wt and TCAIM KI 759 CD8+ T cells. We further minimized the sum of glutamine uptake for wt and TCAIM KI CD8+ 760 T cells together with the sum of regulation parameters to avoid non-physiological futile 761 cycling in the TCA cycle. To optimize the objective function, we used the R package "nloptr" 762 (Steven G. Johnson, The NLopt nonlinear-optimization package, 763 http://github.com/stevengj/nlopt) in a multi start approach to avoid local optima. Further 764 details of the modelling methods are described in (Identification of differential regulation in 765 central carbon metabolism between related cell lines, Roman Rainer, 2020, Humboldt- 766 Universität zu Berlin, http://dx.doi.org/10.18452/22117). 767 768 Plasmids and plasmid constructions. 769 The p TCAIM-EGFP-N1 construct was generated as previously described16 (TCAIM was 770 formerly known as TOAG1). Mito-PAGFP encoding for an eGFP with mitochondrial target 771 sequence was a gift from Richard Youle (Addgene plasmid # 23348; 772 http://n2t.net/addgene:23348; RRID:Addgene_23348)65. 773 For split eGFP fusion protein generation, Tcaim/ Tomm40/ Immt sequences containing a 774 3’- end SapI restriction site inserted into pMA-T vectors as well as c- and n-terminal parts of 775 split eGFP sequences inserted into pMA-RQ and pMK-RQ, respectively, were purchased 776 from Invitrogen. C- and n-terminal split eGFP fragments were ligated into TCAIM, TOM40 or 777 MIC60 coding pMA-T vectors by destroying Tcaim/ Tomm40/ Immt stop codons to create the 778 respective split GFP fusion proteins. Fusion protein coding sequences were excised by 779 BamHI and NotI restriction and inserted into BamHI and NotI restricted pRNA2 vector (gift 780 from Prof. Nina Babel) for expression analysis in eukaryotic systems. 781 782 Co-immunoprecipitation and mass spectrometry analysis. 1.5 x 106 HEK293T cells were 783 seeded in DMEM supplemented with 10 % (vol/ vol) FCS, 100 U/ mL penicillin, 100 mg/ mL 784 streptomycin on a 10 cm tissue culture treated petri dish (Falcon) and 24 h later transfected 785 with either p TCAIM-EGFP-N1 (25 µg) or mito-PAGFP (10 µg) control plasmid DNA using the 786 calcium phosphate technique. In brief, calcium-phosphate-DNA-particle were generated by 787 mixing the plasmid DNA’s with a 250 mM CaCl (Sigma-Aldrich) solution which was then 788 added dropwise to an equal volume of 2 x hepes buffered saline and subsequently incubated 789 for 20 min at room temperature. The precipitated particles were then added dropwise to the 790 cells. After 24 h of incubation at 37 °C and 5 % CO2 the medium was replaced for fresh 791 supplemented DMEM culture medium. The co-immunoprecipitation was then performed 19

bioRxiv preprint doi: https://doi.org/10.1101/2021.04.20.440500; this version posted April 21, 2021. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY-NC-ND 4.0 International license.

792 utilizing the µMACS GFP Isolation Kit (Miltenyi) with protein extracts from whole cell lysates 793 of 2 x 107 transfected HEK293T cells. Deviating to manufacturer’s protocol cells were 794 incubated in lysis buffer supplemented with protease inhibitor (Sigma-Aldrich) in a 1:10 ratio 795 for 1 h on ice. Eluates were separated by SDS-PAGE, stained with PageBlue (Thermo 796 Scientific) and analyzed by liquid chromatography mass spectrometry (LC-MS/MS). Briefly, 797 protein bands in molecular weight range of 40 – 250 kDa were excised and subjected to 798 in-gel tryptic digestion as described previously 66. LC-MS/MS analyses were performed on an 799 Ultimate 3000 RSLCnano system online coupled to an Orbitrap Q Excative Plus mass 800 spectrometer (both Thermo Fisher Scientific). The system comprised a 75 µm i.d. × 250 mm 801 nano LC column (Acclaim PepMap C18, 2 µm; 100 Å; Thermo Fisher Scientific). Mobile 802 phase (A) was 0.1 % formic acid in water and (B) 0.1 % formic acid in 80:20 (v/v) 803 acetonitrile/water. The gradient was 8 – 38 % B in 80 min. The Q ExactivePlus instrument 804 was operated in data dependent mode to automatically switch between full scan MS and 805 MS/MS acquisition. Survey full scan MS spectra (m/ z 350 – 1650) were acquired in the 806 Orbitrap with 70,000 resolution (m/ z 200) after 60 ms accumulation of ions to a 1 x 106 target 807 value. Dynamic exclusion was set to 60 s. The ten most intense multiply charged ions (z ≥ 2) 808 were sequentially isolated and fragmented by higher-energy collisional dissociation (HCD) 809 with a maximal injection time of 120 ms, AGC 5 x 104 and resolution 17,500. 810 For protein identification and relative label-free quantification, MaxQuant software v1.6.0.167 811 with default Andromeda LFQ parameter was used. Spectra were matched to a mus 812 musculus (17,040 reviewed entries, downloaded from .org), a contaminant and decoy 813 database. Acquired MS/MS were searched with parameter as followed: precursor mass 814 tolerance of 10 ppm, fragment tolerance of 0.02 Da, trypsin specificity with a maximum of two 815 missed cleavages, cysteine carbamidomethylation set as fixed and methionine oxidation as 816 variable modification. Identifications were filtered at 1 % False Discovery Rate (FDR) at 817 level. Bioinformatic analysis of the data was performed with Perseus software 818 v1.6.0.2.

819 820 BiFC Assay. 4 x 104 HEK293T cells were seeded in DMEM supplemented with 10 % 821 (vol/ vol) FCS, 100 U/ mL penicillin, 100 mg/ mL streptomycin on a tissue culture treated 24 822 well plate (Falcon) and 24 h later co-transfected with either p TCAIM-EGFPC-RNA2 and 823 pTOM40-EGFPN-RNA2, p TOM40-EGFPN-RNA2 and pMIC60-EGFPC-RNA2 or 824 pTCAIM-EGFPC-RNA2 alone using Lipofectamine 2000 Transfection Reagent (Invitrogen) 825 according to manufacturer’s protocol. A total volume of 500 ng DNA (250 ng per construct) 826 and a final volume of 4 µl Lipofectamine was used for each transfection. After 72 h of 827 incubation at 37 °C and 5 % CO2 cells were trypsinized using Trypsin-EDTA (PAN Biotech) 828 diluted 1:2 in PBS, washed and eGFP expression was examined by flow cytometry using an 829 LSRFortessa with BD FACSDiva software v8.0.2 (BD Biosciences). Data were analyzed with 830 FlowJo v10 software (FlowJo, LLC). 831 832 Transmission electron microscopy. Naïve and polyclonally activated CD8+ T cells were 833 fixed with 0.1 M sodium cacodylate buffer (Serva) supplemented with 2.5 % glutaraldehyde 834 (Serva) for 30 min at room temperature, stored by 4 °C and submitted to the Core Facility for 835 Electron Microscopy (Charité – Universitätsmedizin Berlin). Samples were postfixed with 836 0.1 M sodium cacodylate buffer supplemented with 1 % osmium tetroxide (Science Services) 837 and potassium ferrocyanide (Merck) and embedded in 1 % agarose (Sigma-Aldrich). 838 Embedded samples were dehydrated in a graded ethanol (Merck) series and embedded in 839 epon (Serva). Ultrathin sections of 70 nm were prepared on a Leica Ultracut S (Leica 840 Biosystems). Sections were stained with 4 % uranyl acetate (Serva) and Reynold’s lead 20

bioRxiv preprint doi: https://doi.org/10.1101/2021.04.20.440500; this version posted April 21, 2021. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY-NC-ND 4.0 International license.

841 citrate (Merck). Micrographs were taken with a Zeiss EM 906 transmission electron 842 microscope at 80 kV acceleration (Carl Zeiss) and a slow scan 2k CCD camera (TRS). 843 Mitochondria and cristae were identified manually and the measures were outlined with an 844 optical pen using ImageJ v1.53a software. The roundness factor was calculated as 845 4*area/pi*sqr(major axis) and a maximal distance of 75 nm between mitochondria and ER 846 was defined as MERC. 847 848 Transcription factor binding site prediction. 849 A genome sequence of ~900 bp in close proximity to the transcriptional start site (TSS) and 850 upstream of the translational start (ATG) of the murine Tcaim gene was selected via 851 Ensembl genome browser analyzed using the TFSEARCH online tool 852 (http://diyhpl.us/~bryan/irc/protocol-online/protocol-cache/TFSEARCH.html) with a statistical 853 score threshold above 85 %68. 854 855 Statistical analysis. Quantitative data were represented as scatter dot plots, means, means 856 ± SEM, boxplots with whiskers from min to max or stacked bar charts as indicated in the 857 figure legends. Except for electron microscopy, graph data points of scatter dot plots and 858 boxplots with whiskers from min to max represent biological replicates. Statistical 859 significance was determined as indicated in the figure legends. Datasets were first tested for 860 Gaussian distribution using the D´Agostino and Pearson normality test for sample sizes 861 above five. For smaller sample sizes, the Shapiro-Wilk normality test was applied. If both 862 compared datasets were Gaussian distributed a parametric test was applied, otherwise a 863 non-parametric test was used. P-Values below 0.05 were considered as statistically 864 significant. Statistical analysis was done using GraphPad Prism v7 software (GraphPad 865 Software). 866

21

bioRxiv preprint doi: https://doi.org/10.1101/2021.04.20.440500; this version posted April 21, 2021. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY-NC-ND 4.0 International license.

867 Data availability 868 The RNA-seq data are deposited at the NCBI GEO database under the accession code: 869 GSE167520. 870 871 Acknowledgements 872 We thank the Core Facility for Electron Microscopy of the Charité, (especially Petra Schrade) 873 for support in acquisition of the data. We also thank Dr. T. Borodina and M. Sohn from the 874 Scientific Genomics Platform (Max Delbrück Center for Molecular Medicine) for library 875 preparation and sequencing 876 877 Author contributions 878 C.I., J.S. and B.S. designed the experiments. C.I. performed the experiments with the 879 assistance from C.A., K.V., K.M., K.S., and D.U., analyzed the data and performed the 880 statistical analysis with the exception of: J.L. performed the GC/APCI/MS and analyzed the 881 metabolomics data; A.A.K. performed the immunohistology and analyzed the data; K.J. 882 analyzed the RNA-seq data; R.J.R. constructed the metabolic flux models with the 883 assistance and advice of K.T. and E.K; K.J., K.T.T. and C.G. performed the LC-MS/MS and 884 identification of TCAIM interaction partner. A.P. and S.S. contributed critical reagents and 885 provided experimental help or helped with data analysis. C.M. provided advice to data 886 interpretation. C.I. and B.S. wrote the manuscript. 887 888 Funding Sources 889 This work was supported by the German Research Foundation (DFG) SA 138331 and the 890 EU project Reshape 825392 to B.S. 891 892 Competing interests 893 The authors declare no competing interests. 894 895

22

bioRxiv preprint doi: https://doi.org/10.1101/2021.04.20.440500; this version posted April 21, 2021. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY-NC-ND 4.0 International license.

896 References 897 1. Shyer, J.A., Flavell, R.A. & Bailis, W. Metabolic signaling in T cells. Cell Res. 30, 649- 898 659 (2020). 899 900 2. Chang, C.H. et al. Posttranscriptional control of T cell effector function by aerobic 901 glycolysis. Cell 153, 1239-1251 (2013). 902 903 3. Wang, R.N. et al. The Transcription Factor Myc Controls Metabolic Reprogramming 904 upon T Lymphocyte Activation. Immunity 35, 871-882 (2011). 905 906 4. Finlay, D.K. et al. PDK1 regulation of mTOR and hypoxia-inducible factor 1 integrate 907 metabolism and migration of CD8(+) T cells. J. Exp. Med. 209, 2441-2453 (2012). 908 909 5. Doedens, A.L. et al. Hypoxia-inducible factors enhance the effector responses of 910 CD8(+) T cells to persistent antigen. Nat. Immunol. 14, 1173-U1199 (2013). 911 912 6. Angajala, A. et al. Diverse Roles of Mitochondria in Immune Responses: Novel 913 Insights Into Immuno-Metabolism. Front. Immunol. 9 (2018). 914 915 7. Cao, Y., Rathmell, J.C. & Macintyre, A.N. Metabolic Reprogramming towards Aerobic 916 Glycolysis Correlates with Greater Proliferative Ability and Resistance to Metabolic 917 Inhibition in CD8 versus CD4 T Cells. PLoS One 9 (2014). 918 919 8. Sukumar, M. et al. Mitochondrial Membrane Potential Identifies Cells with Enhanced 920 Stemness for Cellular Therapy. Cell Metab. 23, 63-76 (2016). 921 922 9. Murphy, M.P. & Siegel, R.M. Mitochondrial ROS Fire Up T Cell Activation. Immunity 923 38, 201-202 (2013). 924 925 10. Okoye, I. et al. The protein LEM promotes CD8(+) T cell immunity through effects on 926 mitochondrial respiration. Science 348, 995-1001 (2015). 927 928 11. Sena, L.A. et al. Mitochondria Are Required for Antigen-Specific T Cell Activation 929 through Reactive Oxygen Species Signaling. Immunity 38, 225-236 (2013). 930 931 12. Yerinde, C., Siegmund, B., Glauben, R. & Weidinger, C. Metabolic Control of 932 Epigenetics and Its Role in CD8(+) T Cell Differentiation and Function. Front. 933 Immunol. 10, 2718 (2019). 934 935 13. Buck, M.D. et al. Mitochondrial Dynamics Controls T Cell Fate through Metabolic 936 Programming. Cell 166, 63-76 (2016). 937 938 14. Chao, T., Wang, H. & Ho, P.-C. Mitochondrial Control and Guidance of Cellular 939 Activities of T Cells. Front. Immunol. 8 (2017). 940

23

bioRxiv preprint doi: https://doi.org/10.1101/2021.04.20.440500; this version posted April 21, 2021. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY-NC-ND 4.0 International license.

941 15. Schumann, J. et al. The Mitochondrial Protein TCAIM Regulates Activation of T Cells 942 and Thereby Promotes Tolerance Induction of Allogeneic Transplants. Am. J. 943 Transplant. 14, 2723-2735 (2014). 944 945 16. Keeren, K. et al. Expression of Tolerance Associated Gene-1, a Mitochondrial Protein 946 Inhibiting T Cell Activation, Can Be Used to Predict Response to Immune Modulating 947 Therapies. J. Immunol. 183, 4077-4087 (2009). 948 949 17. Klein-Hessling, S. et al. NFATc1 controls the cytotoxicity of CD8(+) T cells. Nature 950 Communications 8 (2017). 951 952 18. Hogan, P.G., Chen, L., Nardone, J. & Rao, A. Transcriptional regulation by calcium, 953 calcineurin, and NFAT. Genes Dev. 17, 2205-2232 (2003). 954 955 19. Dimeloe, S., Burgener, A.V., Grahlert, J. & Hess, C. T-cell metabolism governing 956 activation, proliferation and differentiation; a modular view. Immunology 150, 35-44 957 (2017). 958 959 20. Zorova, L.D. et al. Functional Significance of the Mitochondrial Membrane Potential. 960 Biochem Mosc Suppl S 12, 20-26 (2018). 961 962 21. Phan, A.T. & Goldrath, A.W. Hypoxia-inducible factors regulate T cell metabolism and 963 function. Mol. Immunol. 68, 527-535 (2015). 964 965 22. Pålsson-McDermott, E.M. & O’Neill, L.A.J. Targeting immunometabolism as an anti- 966 inflammatory strategy. Cell Res. 30, 300-314 (2020). 967 968 23. Isaacs, J.S. et al. HIF overexpression correlates with biallelic loss of fumarate 969 hydratase in renal cancer: Novel role of fumarate in regulation of HIF stability. Cancer 970 Cell 8, 143-153 (2005). 971 972 24. Fischer, M. et al. Early effector maturation of naïve human CD8+ T cells requires 973 mitochondrial biogenesis. Eur. J. Immunol. 48, 1632-1643 (2018). 974 975 25. Konjar, Š. et al. Mitochondria maintain controlled activation state of epithelial-resident 976 T lymphocytes. Science Immunology 3, eaan2543 (2018). 977 978 26. López-Crisosto, C. et al. ER-to-mitochondria miscommunication and metabolic 979 diseases. Biochimica et Biophysica Acta (BBA) - Molecular Basis of Disease 1852, 980 2096-2105 (2015). 981 982 27. Martinvalet, D. The role of the mitochondria and the endoplasmic reticulum contact 983 sites in the development of the immune responses. Cell Death Dis. 9, 336 (2018). 984 985 28. Modi, S. et al. Miro clusters regulate ER-mitochondria contact sites and link cristae 986 organization to the mitochondrial transport machinery. Nature Communications 10 987 (2019).

24

bioRxiv preprint doi: https://doi.org/10.1101/2021.04.20.440500; this version posted April 21, 2021. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY-NC-ND 4.0 International license.

988 989 29. von der Malsburg, K. et al. Dual Role of Mitofilin in Mitochondrial Membrane 990 Organization and Protein Biogenesis. Dev. Cell 21, 694-707 (2011). 991 992 30. Chen, H.W., Heiniger, H.J. & Kandutsch, A.A. Relationship between Sterol Synthesis 993 and DNA-Synthesis in Phytohemagglutinin-Stimulated Mouse Lymphocytes. Proc. 994 Natl. Acad. Sci. U. S. A. 72, 1950-1954 (1975). 995 996 31. Chakrabarti, R. & Engleman, E.G. Interrelationships between Mevalonate Metabolism 997 and the Mitogenic Signaling Pathway in Lymphocyte-T Proliferation. J. Biol. Chem. 998 266, 12216-12222 (1991). 999 1000 32. Kidani, Y. et al. Sterol regulatory element-binding proteins are essential for the 1001 metabolic programming of effector T cells and adaptive immunity. Nat. Immunol. 14, 1002 489-+ (2013). 1003 1004 33. Yang, W. et al. Potentiating the antitumour response of CD8(+) T cells by modulating 1005 cholesterol metabolism. Nature 531, 651-+ (2016). 1006 1007 34. Sala-Vila, A. et al. Interplay between hepatic mitochondria-associated membranes, 1008 lipid metabolism and caveolin-1 in mice. Sci. Rep. 6 (2016). 1009 1010 35. Csordas, G. et al. Imaging Interorganelle Contacts and Local Calcium Dynamics at 1011 the ER-Mitochondrial Interface. Mol. Cell 39, 121-132 (2010). 1012 1013 36. De Vos, K.J. et al. VAPB interacts with the mitochondrial protein PTPIP51 to regulate 1014 calcium homeostasis. Hum. Mol. Genet. 21, 1299-1311 (2012). 1015 1016 37. Rizzuto, R. et al. Close contacts with the endoplasmic reticulum as determinants of 1017 mitochondrial Ca2+ responses. Science 280, 1763-1766 (1998). 1018 1019 38. Szabadkai, G. et al. Chaperone-mediated coupling of endoplasmic reticulum and 1020 mitochondrial Ca2+ channels. J. Cell Biol. 175, 901-911 (2006). 1021 1022 39. Anavi, S., Hahn-Obercyger, M., Madar, Z. & Tirosh, O. Mechanism for HIF-1 1023 activation by cholesterol under normoxia: A redox signaling pathway for liver damage. 1024 Free Radic. Biol. Med. 71, 61-69 (2014). 1025 1026 40. Daley, S.R. et al. Rasgrp1 mutation increases naive T-cell CD44 expression and 1027 drives mTOR-dependent accumulation of Helios(+) T cells and autoantibodies. Elife 2 1028 (2013). 1029 1030 41. Dodd, K.M., Yang, J., Shen, M.H., Sampson, J.R. & Tee, A.R. mTORC1 drives HIF-1 1031 alpha and VEGF-A signalling via multiple mechanisms involving 4E-BP1, S6K1 and 1032 STAT3. Oncogene 34, 2239-2250 (2015). 1033

25

bioRxiv preprint doi: https://doi.org/10.1101/2021.04.20.440500; this version posted April 21, 2021. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY-NC-ND 4.0 International license.

1034 42. Eid, W. et al. mTORC1 activates SREBP-2 by suppressing cholesterol trafficking to 1035 lysosomes in mammalian cells. Proc. Natl. Acad. Sci. U. S. A. 114, 7999-8004 (2017). 1036 1037 43. Porstmann, T. et al. SREBP activity is regulated by mTORC1 and contributes to Akt- 1038 dependent cell growth. Cell Metab. 8, 224-236 (2008). 1039 1040 44. Yang, K. et al. T Cell Exit from Quiescence and Differentiation into Th2 Cells Depend 1041 on Raptor-mTORC1-Mediated Metabolic Reprogramming. Immunity 39, 1043-1056 1042 (2013). 1043 1044 45. Schieke, S.M. et al. The mammalian target of rapamycin (mTOR) pathway regulates 1045 mitochondrial oxygen consumption and oxidative capacity. J. Biol. Chem. 281, 27643- 1046 27652 (2006). 1047 1048 46. Head, S.A. et al. Antifungal drug itraconazole targets VDAC1 to modulate the 1049 AMPK/mTOR signaling axis in endothelial cells. Proc. Natl. Acad. Sci. U. S. A. 112, 1050 E7276-E7285 (2015). 1051 1052 47. Camara, A.K.S., Zhou, Y.F., Wen, P.C., Tajkhorshid, E. & Kwok, W.M. Mitochondrial 1053 VDAC1: A Key Gatekeeper as Potential Therapeutic Target. Front. Physiol. 8 (2017). 1054 1055 48. Ramanathan, A. & Schreiber, S.L. Direct control of mitochondrial function by mTOR. 1056 Proc. Natl. Acad. Sci. U. S. A. 106, 22229-22232 (2009). 1057 1058 49. Sood, A. et al. A Mitofusin-2-dependent inactivating cleavage of Opa1 links changes 1059 in mitochondria cristae and ER contacts in the postprandial liver. Proc. Natl. Acad. 1060 Sci. U. S. A. 111, 16017-16022 (2014). 1061 1062 50. Bantug, G.R. et al. Mitochondria-Endoplasmic Reticulum Contact Sites Function as 1063 Immunometabolic Hubs that Orchestrate the Rapid Recall Response of Memory 1064 CD8+ T Cells. Immunity 48, 542-555.e546 (2018). 1065 1066 References Methods 1067 51. Ocker, L. et al. Hypericin and its radio iodinated derivatives - A novel combined 1068 approach for the treatment of pediatric alveolar rhabdomyosarcoma cells in vitro. 1069 Photodiagnosis Photodyn. Ther. 29 (2020). 1070 1071 52. Kopka, J. et al. [email protected]: the Golm Metabolome Database. Bioinformatics 21, 1072 1635-1638 (2005). 1073 1074 53. Jaeger, C., Hoffmann, F., Schmitt, C.A. & Lisec, J. Automated Annotation and 1075 Evaluation of In-Source Mass Spectra in GC/Atmospheric Pressure Chemical 1076 Ionization-MS-Based Metabolomics. Anal. Chem. 88, 9386-9390 (2016). 1077 1078 54. Jaeger, C. & Lisec, J. Statistical and Multivariate Analysis of MS-Based Plant 1079 Metabolomics Data. Methods Mol. Biol. 1778, 285-296 (2018).

26

bioRxiv preprint doi: https://doi.org/10.1101/2021.04.20.440500; this version posted April 21, 2021. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY-NC-ND 4.0 International license.

1080 1081 55. Schubert, M., Lindgreen, S. & Orlando, L. AdapterRemoval v2: rapid adapter 1082 trimming, identification, and read merging. BMC Res. Notes 9, 88 (2016). 1083 1084 56. Kim, D. et al. TopHat2: accurate alignment of transcriptomes in the presence of 1085 insertions, deletions and gene fusions. Genome Biol. 14, R36 (2013). 1086 1087 57. Langmead, B. & Salzberg, S.L. Fast gapped-read alignment with Bowtie 2. Nat 1088 Methods 9, 357-359 (2012). 1089 1090 58. Liao, Y., Smyth, G.K. & Shi, W. The R package Rsubread is easier, faster, cheaper 1091 and better for alignment and quantification of RNA sequencing reads. Nucleic Acids 1092 Res. 47 (2019). 1093 1094 59. Love, M.I., Huber, W. & Anders, S. Moderated estimation of fold change and 1095 dispersion for RNA-seq data with DESeq2. Genome Biol. 15 (2014). 1096 1097 60. Fresno, C. & Fernandez, E.A. RDAVIDWebService: a versatile R interface to DAVID. 1098 Bioinformatics 29, 2810-2811 (2013). 1099 1100 61. Yu, G.C., Wang, L.G., Han, Y.Y. & He, Q.Y. clusterProfiler: an R Package for 1101 Comparing Biological Themes Among Gene Clusters. Omics-a Journal of Integrative 1102 Biology 16, 284-287 (2012). 1103 1104 62. Walter, W., Sanchez-Cabo, F. & Ricote, M. GOplot: an R package for visually 1105 combining expression data with functional analysis. Bioinformatics 31, 2912-2914 1106 (2015). 1107 1108 63. Thiele, I. et al. A community-driven global reconstruction of human metabolism. Nat. 1109 Biotechnol. 31, 419-+ (2013). 1110 1111 64. Lisec, J., Hoffmann, F., Schmitt, C. & Jaeger, C. Extending the Dynamic Range in 1112 Metabolomics Experiments by Automatic Correction of Peaks Exceeding the 1113 Detection Limit. Anal. Chem. 88, 7487-7492 (2016). 1114 1115 65. Karbowski, M. et al. Quantitation of mitochondrial dynamics by photolabeling of 1116 individual organelles shows that mitochondrial fusion is blocked during the Bax 1117 activation phase of . J. Cell Biol. 164, 493-499 (2004). 1118 1119 66. Lehmann, A., Niewienda, A., Jechow, K., Janek, K. & Enenkel, C. Ecm29 Fulfils 1120 Quality Control Functions in Proteasome Assembly. Mol. Cell 38, 879-888 (2010). 1121 1122 67. Cox, J. et al. Accurate Proteome-wide Label-free Quantification by Delayed 1123 Normalization and Maximal Peptide Ratio Extraction, Termed MaxLFQ. Mol. Cell. 1124 Proteomics 13, 2513-2526 (2014). 1125

27

bioRxiv preprint doi: https://doi.org/10.1101/2021.04.20.440500; this version posted April 21, 2021. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY-NC-ND 4.0 International license.

1126 68. Heinemeyer, T. et al. Databases on transcriptional regulation: TRANSFAC, TRRD 1127 and COMPEL. Nucleic Acids Res. 26, 362-367 (1998). 1128 1129

28

bioRxiv preprint doi: https://doi.org/10.1101/2021.04.20.440500; this version posted April 21, 2021. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY-NC-ND 4.0 International license.

a b G lu c o s e P y ru v a te L a c ta t e TCAIM KI wt (KI) TCAIM KO wt (KO) 4 .9 4 .0 6 .2 17.3 16.7 49.3 26.2 51.4 25.8 52.6 31.1 4 .7 3 .5 6 .0 In te n s it y

10 4 .5 3 .0 5 .8

4 .3 2 .5 5 .6 59.7 21.0 19.3 14.1 CD44 4 .1 2 .0 5 .4 CD62L

N o r m . L o g K I w t K I w t K I w t

Glycerol-3- Sphingosine A la n in e c 0 .0 0 4 5 0 .0 4 0 5 0 .0 0 9 6 – + p h o s . 1 0 0 + 1 0 0 6 0 5 .0 4 .0 7 .0 8 0 8 0 4 5 3 .5 6 .0 4 .8 6 0 6 0 C D 6 2 L In te n s it y C D 6 2 L C D 6 2 L + – + 3 0 10 3 .0 5 .0 4 0 4 0 4 .6 1 5 2 .5 4 .0 2 0 2 0 0 0 0 % C D 4 4 % C D 4 4 4 .4 2 .0 3 .0 % C D 4 4 K I w t K O w t K I w t K O w t K I w t K O w t N o r m . L o g K I w t K I w t K I w t

S e rin e P ro lin e A s p a r ta t e d e 0 .0 0 3 0 0 .0 1 2 3 1 0 0 1 0 0 6 .0 8 .0 3 .7

+ 8 0 + 8 0 6 .0 3 .5 4 .0 6 0 6 0 In te n s it y

10 4 .0 3 .3 4 0 4 0 2 .0 % C D 6 9 2 .0 3 .1 2 0 % C D 2 5 2 0 0 0 0 .0 0 .0 2 .9 K I w t K O w t K I w t K O w t N o r m . L o g K I w t K I w t K I w t f g 2 1 2 0 .0 0 0 6 8 0 KI 1 0 0 3 w t ( K I) 8 0 6 0 + KO 6 0 4 0 w t ( K O ) 1 4 0 3

2 0 % K i- 6 7 0 0 2 0 Norm. to Mode toNorm. 0 0 Proliferation Dye % Proliferated cells0 1 2 3 4 5 6 K I w t K O w t # cell divisions h i j TCAIM KO wt (KO) 0 .0 2 2 3 0 .0 2 0 6 1 .5 1 .5 

1 .0 1 .0

0 .5 0 .5 (rel. to wt) (rel. to wt) geoMFI IFN- 0 .0 geoMFI GZMB 0 .0 KIKO KIKO k

1 5 0

1 0 0

T cells (105 hpf) 0 +

0 K O w t # IE C D 8 bioRxiv preprint doi: https://doi.org/10.1101/2021.04.20.440500; this version posted April 21, 2021. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY-NC-ND 4.0 International license.

Fig. 1 │TCAIM KI alters the CD8+ T cell metabolic program and inhibits their proliferation and effector differentiation. (a) Non-targeted metabolite analysis of 60 h polyclonally activated TCAIM KI or wt CD8+ T cells (n = 3). Representative contour plots (b) and quantification of CD62L/ CD44 (c), CD69 (d) and CD25 (e) protein surface expression of 72 h activated TCAIM KI, KO and respective wt CD8 T cells (n = 9). Cells were polyclonally activated with either 5 µg/ ml αCD3 and 2 µg/ ml αCD28 or 2 µg/ ml αCD3 and 1 µg/ ml αCD28. ( f) Representative histogram overlays and quantification of in vitro⁺proliferation of 48 h polyclonally activated TCAIM KI or wt CD8+ T cells co-cultured with B cells (n = 6). (g) Ki-67 protein expression of 48 h polyclonally activated TCAIM KI, KO and respective wt CD8 T cells (n = 6). Intracellular IFN-γ (h) and GZMB (i) protein expression levels relative to wt controls of 48 h polyclonally activated TCAIM KI or KO CD8 T cells co-cultured with B cells (n = 6). (j) Representative⁺ colon sections stained for CD3 (green), CD4 (blue) and CD8 (red) collected from T cell specific TCAIM KO mice or wt littermates. White arrowheads⁺ indicate epithelia infiltrating CD3+/ CD8+ T cells. Scale bars: 25 µm. (k) Absolute number of intraepithelial (IE) CD3+/ CD8+ T cells per 10 high power fields (hpf) in immunohistochemically stained colon sections (n = 2). Cells had been stimulated with 5 µg/ ml plate-bound αCD3 and 2 µg/ ml soluble αCD28 in a density of 6 x 105 cells/ 200 µl culture medium (a,f-i). Quantitative data represents means (a), boxplots with whiskers from min to max (c-e,g-i), means ± SEM (f) or scatter dot plots (k). Data were analyzed by one-tailed, paired Student’s t-test (c-e), one-tailed, unpaired Student’s t-test (g-h) or one-tailed Mann-Whitney test (i). geoMFI , geometric mean fluorescence intensity. bioRxiv preprint doi: https://doi.org/10.1101/2021.04.20.440500; this version posted April 21, 2021. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY-NC-ND 4.0 International license.

a b Murine Tcaimp (D9Ertd402e promoter) 866 bp 0.0466 0.0061 0 .0 6 TSS ATG +1 start 0 .0 4

(NFAT) (NFAT) (CREB) (C/EBP) (NFAT) (C/EBP) (C/EBP) m R N A

-172/ -168 -158/ -154 -72/ -60 +58/ +69 +131/ +135 +259/ +270 +395/ +407 0 .0 2 T c a im 0 .0 0 c (rel. to Ubc mRNA) – – CsA 0.0452 0.0179 0.0040 0.0247 1 0 0 24 h 48 h 1 0 0 1 0 - 1 1 0 - 1 1 0 - 2 1 0 - 2 1 0 - 3

m R N A - 3 m R N A 1 0 - 4  1 0 - 5 - 4 Il- 2 1 0 1 0 Ifn - 1 0 - 6 1 0 - 5 (rel. to Ubc mRNA) (rel. to Ubc mRNA) – – CsA – – CsA 24 h 48 h 24 h 48 h

Fig. 2 │ NFAT signaling regulates Tcaim gene expression. (a) Schematic representation of the murine D9Ertd402e promoter (Tcaimp) with predicted transcription factors binding sites displayed in brackets. Distance down- or upstream form the transcription start site (TSS) is indicated with - or +, respectively. Relative Tcaim (b) or Il-2 and Ifn-γ (c) mRNA expression analyzed by real-time qRT-PCR of 24 - 48 h polyclonally activated C57BL/6 CD8+ T cells upon treatment with the calcineurin inhibitor Cyclosporine A (CsA) or DMSO control (n = 5). Cells had been stimulated with 5 µg/ ml plate-bound αCD3 and 2 µg/ ml soluble αCD28 at a density of 6 x 105 cells/ 200 µl culture medium. Gene expression was normalized to Ubc expression. Quantitative data represents boxplots with whiskers from min to max. Data were analyzed by one-tailed, unpaired Mann-Whitney test (Ifn-γ 24 vs. 48 h DMSO) or one-tailed, unpaired Student’s t-test (others). bioRxiv preprint doi: https://doi.org/10.1101/2021.04.20.440500; this version posted April 21, 2021. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY-NC-ND 4.0 International license.

a b Glycolysis

Aldoa Aldoc Eno1 Eno2 Eno3 Gapdh Gpi1 Hk1 Hk2 Hk3 Ldha Ldhb Ldhd Pfkl Pfkm Pfkp Pgam1 Pgam2 Pgk1 Pklr Pkm2 Slc2a1 Slc2a3 Tpi1 0 h 24 h 48 h

TCA cycle

Aco1 Aco2 Cs Dld Dlst Fh1 Idh1 Idh2 Idh3a Idh3b Idh3g Mdh2 Ogdh Ogdhl Sdha Sdhb Sdhc Sdhd Sucla2 Suclg1 Suclg2 Me2 0 h 24 h 48 h

Glutaminolysis

Gls Gls2 Glud1 Got1 Got2 Gpt 0 h 24 h 48 h

Transcription factors

Myc Hif1a Hif2a Hif3a 0 h 24 h 48 h

Row Z-Score 0 2-2 bioRxiv preprint doi: https://doi.org/10.1101/2021.04.20.440500; this version posted April 21, 2021. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY-NC-ND 4.0 International license.

Fig. 3 │ Reduced enzyme gene expression and metabolic flux rates in TCAIM KI CD8+ T cells. (a) Heatmap displaying normalized expression values (z-score) of selected genes encoding for glycolysis, tricarboxylic acid cycle and glutaminolysis controlling enzymes or transcription factors in naïve and 24 to 48 h polyclonally activated TCAIM KI (blue), KO (red) or wt (grey) CD8+ T cells (KI and wt: n = 3; KO: n = 2). Highlighted genes indicate those being differentially expressed between 48 h activated TCAIM KI and wt CD8+ T cells. P-values have been adjusted to either 0.05 (bold blue) or 0.3 (bold black). Expression of protein-coding genes was determined by normalization and variance-stabilized transformation of RNA-seq raw count data. (b) Mathematical metabolic flux modeling for 60 h polyclonally activated TCAIM KI or wt CD8+ T cells. Via mass spectrometry, measured metabolites are highlighted in purple boxes while non-measured metabolites are framed in grey. The flux distributions in mM / h according to the calculated enzyme activity of TCAIM KI and wt CD8+ T cells are displayed as bar graphs for each enzyme, respectively (wt: grey bar; KI: blue bar). Cells had been stimulated with 5 µg/ ml plate-bound αCD3 and 2 µg/ ml soluble αCD28 at a density of 6 x 105 cells/ 200 µl culture medium. bioRxiv preprint doi: https://doi.org/10.1101/2021.04.20.440500; this version posted April 21, 2021. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY-NC-ND 4.0 International license.

a b h h

0 0 24 24 1 48 48 72 72 96 96 Norm. to Mode toNorm. TMRE 0 5 0 0 * *** KI 4 0 0 w t ( K I) KO 3 0 0 w t ( K O ) -1 2 0 0 (RNA foldchange)(RNA (x 1 0 0 0 )

2 1 0 0

geoMFI TMRE 0 Log

-2 0 2 4 4 8 7 2 9 6

tim e ( h ) c 0 h -3 0.0278 0.0601 2 0 -3 -2 -1 0 1 1 5 Log 2 (Regularization parameter) d 1 0

hocre (KI) hecre (KI) wtcre (KI) (x 1 0 0 0 ) 5

0 h 0.05 3.70 4.33 geoMFI TMRE 0 K I w t K O w t e 0 h 7 2 h 0.0465 1 0 0 1 0 0 24 h 50.9 60.6 75.3 8 0 8 0 high

 6 0 6 0 4 0 4 0 2 0 2 0 % H IF 1 CD8 0 0 HIF1α

h o c reh (K e c I)rew (K tc I) re (K I) h o c reh (K e c I) re w(K tc I) re (K I) Fig. 4 │ Reduced ΔΨm and HIF1α protein expression impairs metabolic activity in TCAIM KI CD8+ T cells. (a) Comparison of differentially expressed genes of 48 h polyclonally activated TCAIM + KI or wt CD8 T cells from RNA-seq analysis (log2 fold change) to regularization parameters introduced to describe the deviating TCAIM KI from the wt metabolic fluxes (log2). Only differentially expressed genes with adjusted p-values of ≤ 0.05 (black) or ≤ 0.3 (purple) are included. Enzymatic fluxes are described with either forward (kf) or backward (kb) flux rates. (b) Representative histogram overlays and quantification of TMRE staining for ΔΨm of naïve and polyclonally activated TCAIM KI, KO and respective wt CD8+ T cells co-cultured with B cells (n = 5). Significant differences between TCAIM KI or KO and their wt controls are indicated with blue or red asterisks, respectively. (c) TMRE staining for baseline ΔΨm of naïve TCAIM KI, KO and respective wt CD8+ T cells co-cultured with B cells (n = 5). Representative contour plots (d) and quantification (e) of HIF1α expression of naïve and 72 h polyclonally activated homozygous (hocre), heterozygous (hecre) TCAIM KI or wild type (wtcre) CD8 T cells (n = 6). Cells had been stimulated with 5 µg/ ml plate-bound αCD3 and 2 µg/ ml soluble αCD28 at a density of 6 x 105 cells/ 200 µl culture medium (a). Quantitative flow cytometry data represents⁺ means ± SEM (b) or boxplots with whiskers from min to max (c,e). Data were analyzed according to Gaussian distribution of the compared sample sets and for each genotype with its respective wt control using (b) Kruskal-Wallis test with Dunn’s multiple comparisons test (KI vs. wt) or one-way ANOVA (KO vs. wt) and subsequent post hoc one-tailed Mann-Whitney test (KI vs. wt 0 h) or one-tailed, unpaired Student’s t-test (others), (c) using one-tailed Mann-Whitney test (KI vs wt) or one-tailed, unpaired Student’s t-test (KO vs wt) or (e) Kruskal-Wallis test with Dunn’s multiple comparisons test and subsequent post hoc one-tailed, unpaired Mann-Whitney test. * p < 0.05, ** p < 0.01. geoMFI , geometric mean fluorescence intensity. bioRxiv preprint doi: https://doi.org/10.1101/2021.04.20.440500; this version posted April 21, 2021. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY-NC-ND 4.0 International license.

a b Row Z-Score

20- 2

Z-Score upregulated downregulated 0- 3 4.4

ID Description GO:0001816 Cytokine production GO:0006952 Defense response GO:0009605 Response to external stimulus GO:0016126 sterol biosynthetic process Regulation of cell differentiation 0 h 24 h 48 h GO:0045595 GO:0048585 Negative regulation of response to stimulus GO:0045597 Positive regulation of cell differentiation c GO:0001818 Negative regulation of cytokine production P m v k M vd F d p s

1 3 1 3 1 3 KI 1 2 1 2 1 2

E x p r . w t ( K I) 2 1 1 1 1 1 1 KO

1 0 1 0 1 0

9 9 9 N o r m . L o g 0 2 4 4 8 0 2 4 4 8 0 2 4 4 8

tim e ( h ) d F d ft 1 S q le L s s C y p 5 1 M s m o 1 N s d h l

1 4 1 4 1 4 1 6 1 4 1 3

1 3 1 3 1 3 1 3 1 2 E x p r .

2 1 4 1 2 1 2 1 2 1 2 1 1 1 2 1 1 1 1 1 1 1 1 1 0

1 0 1 0 1 0 1 0 1 0 9 N o r m . L o g 0 2 4 4 8 0 2 4 4 8 0 2 4 4 8 0 2 4 4 8 0 2 4 4 8 0 2 4 4 8 time (h) bioRxiv preprint doi: https://doi.org/10.1101/2021.04.20.440500; this version posted April 21, 2021. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY-NC-ND 4.0 International license.

Fig. 5 │ TCAIM inhibits activation-induced expression of cholesterol biosynthesis genes. (a) Non-supervised hierarchical clustering analysis of 1000 genes with the highest variances across naïve and polyclonally activated TCAIM KI (blue), KO (red) or wt (grey) CD8+ T cells (KI and wt: n = 3; KO: n = 2). Normalized expression values (z-scores) are depicted. Expression of protein-coding genes was determined by normalization and variance-stabilized transformation of RNA-seq raw count data. (b) DAVID functional gene ontology (GO) enrichment analysis within the sets of differentially expressed genes between 48 h polyclonally activated TCAIM KI or wt CD8+ T cells (n = 3). GO circle plot displaying representative GO terms selected out of the top 25 GO biological processes. The outer ring shows the log2 fold change (FC) of the assigned genes from higher to lower (outer to inner layer) for each term. Red circles display genes being upregulated and blue ones being downregulated. The inner ring shows the z-Score [(#genesup - #genesdown) / (#genesup + #genesdown)] corresponding to the blue red gradient. The width of the inner ring bars indicate the significance of each GO term in relation to each other – the wider the bar, the smaller the p value (log10-adjusted p-value < 0.05). mRNA expression of enzymes controlling the mevalonate pathway (c) and cholesterol biosynthesis (d) of naïve and polyclonally activated TCAIM KI, KO or wt CD8+ T cells (KI and wt: n = 3; KO: n = 2).

RNA-seq raw count data of protein-coding genes were normalized and log2-transformed after adding a pseudocount to each value. Cells had been stimulated with 5 µg/ ml plate-bound αCD3 and 2 µg/ ml soluble αCD28 at a density of 6 x 105 cells/ 200 µl culture medium. Quantitative data represent means ± SEM (c-d). bioRxiv preprint doi: https://doi.org/10.1101/2021.04.20.440500; this version posted April 21, 2021. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY-NC-ND 4.0 International license.

a b hocre (KI) hecre (KI) wtcre (KI) 0.0272

9.72 12.3 14.7 19.9 24.0 32.1 – 0 µM 1 0 0 0.0089 n .s . 8 0 C D 4 4

+ 6 0 4 0 71.1 61.4 39.6 2 0 9.92 12.8 19.4 34.5 19.6 40.0 2 µM 0 % C D 6 2 L – – – Chol. hocre (KI) hecre (KI) wtcre (KI)

0.0208 70.4 43.5 37.3

+ 1 0 0 0.0134 n .s . 10 µM 10.8 35.2 14.8 46.1 15.3 41.9 8 0 C D 4 4

+ 6 0 4 0

51.0 37.4 40.2 2 0 0 40 µM 13.7 74.7 25.5 65.5 40.0 48.3 % C D 6 2 L – – – Chol. hocre (KI) hecre (KI) wtcre (KI)

+ 5 0

CD44 10.1 7.45 8.88 4 0 CD62L C D 4 4

– 3 0 c 2 0 0.0403 1 0 0.0291 n .s . 1 0 0 0 % C D 6 2 L 8 0 – – – Chol. high 6 0 hocre (KI) hecre (KI) wtcre (KI) 4 0

% H if1 a 2 0 0 – – – Chol. hocre (KI) hecre (KI) wtcre (KI)

0.0078 d 0 µM 2 µM 10 µM 40 µM e 1 0 0 0.0167 n .s . 0 0 0 1 8 0 0 1 6 0 1 1 2 4 0 2 2 2 2 0

Norm. to Mode toNorm. 0 Proliferation Dye % Proliferated cells – – – Chol. hocre (KI) hecre (KI) wtcre (KI) hocre (KI) hecre (KI) wtcre (KI)

Fig. 6 │ Exogenous cholesterol restores activation and proliferation of TCAIM KI CD8+ T cells. Representative contour plots (a) with quantification of CD62L / CD44 surface protein expression (b), intracellular HIF1α protein expression (c) and representative histogram overlays (d) with quantification of in vitro proliferation (e) of 72 h polyclonally activated homozygous (hocre), heterozygous (hecre) TCAIM KI or wild type (wtcre) CD8 T cells (n = 5). Cells have been treated with increasing doses of Cholesterol or have been left untreated as indicated. Quantitative data represents boxplots with whiskers from min to max and were⁺ analyzed by one way ANOVA with post hoc one-tailed, unpaired Student’s t-test. n.s., not significant (p > 0.05). bioRxiv preprint doi: https://doi.org/10.1101/2021.04.20.440500; this version posted April 21, 2021. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY-NC-ND 4.0 International license.

a TCAIM KI wt (KI) TCAIM KO wt (KO) 0 h

24 h

48 h

b KI wt KO wt c **** **** * **** 0 h 2 4 h 4 8 h 100 * ** **** **** * **** 6 0.0254 90 5 80 4 70 3 60 2

50 # MERCs / 1 with ER / cell 0 mitochondrion 0 % Mito. in contact K I w t K O w t K I w t K O w t h h h h h h h h h h h h KI wt KO wt 0 4 8 0 4 8 0 4 8 0 4 8 2 4 2 4 2 4 2 4 d 0 h 2 4 h 4 8 h

1 0 0 0.0087 8 0 6 0 4 0 2 0

mito. perimeter0 % MERCs length / K I w t K O w t K I w t K O w t K I w t K O w t

Fig. 7 │TCAIM KI delays MERC formation. (a) Zoomed regions of representative transmission electron microscopy images of naïve (upper row) and 24 to 48 h (middle and bottom row, respectively) polyclonally activated TCAIM KI, KO or respective wt CD8+ T cells. White arrowheads indicate mitochondria-ER interaction sites. Scale bars: 500 nm. Digitally magnification: 2.35 x. (b-d) Quantitative plots of transmission electron microscopy image analysis of naïve and polyclonally activated TCAIM KI, KO and respective wt CD8+ T cells. (b) Frequency of mitochondria being in contact with the ER per cell (n = 30). (c) Absolute number of mitochondria-ER contact sites (MERCs) per mitochondrion. (d) Proportional share of total MERCs length’s to mitochondrial perimeter per mitochondrion. For (c) and (d) a minimum of seven cells with a minimum of 30 mitochondria being in contact with the ER have been analyzed. If there were less than 30 mitochondria with ER contact sites the number of cells being analyzed was increased to a maximum of up to 9 cells. bioRxiv preprint doi: https://doi.org/10.1101/2021.04.20.440500; this version posted April 21, 2021. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY-NC-ND 4.0 International license.

a b TOM40 T c a im V d a c 2 Rmdn3 1 4 .5 12.5 1 4 12.0 5 E x p r .

TCAIM 2 1 4 .0 1 2 ODO1 11.5 1 3 .5 1 0 11.0 4 SAM50 8 1 3 .0 PP6R3 N o r m . L o g 10.5 RMD3 0 2 4 4 8 0 2 4 4 8 0 24 48 DJC11 VDAC2 tim e ( h ) 3 MIC60 KBP -value t-test)-value EFTU

(P TRAP1 c S a m m 5 0 I m m t 10 CISY STML2

og 2 1 4 .0 1 4 .5 KI -L w t ( K I) 1 4 .0 E x p r .

2 1 3 .5 KO 1 1 3 .5 1 3 .0 1 3 .0

1 2 .5 1 2 .5 0 N o r m . L o g 0 2 4 4 8 0 2 4 4 8

-12 -10 -8 -6 -4 -2 0 2 4 time (h)

Log 2 Fold change (LFQ mito-eGFP / Tcaim-eGFP) d TCAIM + TOM40 + TCAIM- TOM40 MIC60 eGFPc 0.31 0.28 0.0 FSC-A

eGFP

Fig. 8 │ TCAIM interacts with MERC promoting proteins. (a) GFP co-immunoprecipitates of cell lysates from HEK293T cells transfected with pTCAIM-EGFP-N1 or mito-PAGFP control plasmid were analyzed for TCAIM interaction partner by LC -MS/MS. Volcano plot showing 14 proteins (FDR 0.05, S0 0.7) being significantly enriched only in pTCAIM-EGFP-N1 cell lysate fractions (n = 3). (b-c) mRNA expression of selected TCAIM interaction partners of naïve and polyclonally activated TCAIM KI, KO or wt CD8+ T cells (KI and wt: n = 3; KO: n = 2). RNA-seq raw count data of protein-coding

genes were normalized and log2-transformed after adding a pseudocount to each value. Cells had been stimulated with 5 µg/ ml plate-bound αCD3 and 2 µg/ ml soluble αCD28 at a density of 6 x 105 cells/ 200 µl culture medium. (d) Representative dot plots from one out of two BiFC assay experiments for validation of TCAIM interaction with TOM40 measured by flow cytometry. As indicated HEK293T cells were transfected with plasmids encoding for the split eGFP fusion proteins TCAIM-eGFPc + TOM40-eGFPn, TOM40-eGFPn + MIC60-eGFPc as positive control or TCAIM-eGFPc alone as negative control. Quantitative data represents means ± SEM (b-c). bioRxiv preprint doi: https://doi.org/10.1101/2021.04.20.440500; this version posted April 21, 2021. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY-NC-ND 4.0 International license. a b Glycolysis pathway TCA cycle pathway

G lu c o s e - Glyceric acid- Dihydroxyace- Phosphoenol- 6 - p h o s . 3 - p h o s . to n e p h o s . p y ru v a te C itra te  -Ketoglutarate F u m a r a te M a la te

3 .4 3 .4 3 .1 3 .2 5 .5 4 .6 5 .0 4 .9

3 .2 3 .2 2 .9 3 .0 5 .4 4 .5 4 .7 4 .5 In te n s it y

10 3 .0 3 .0 2 .7 2 .8 5 .3 4 .4 4 .5 4 .0 2 .8 2 .8 2 .5 2 .6 5 .2 4 .3 4 .3 2 .3 2 .6 2 .6 2 .4 5 .1 4 .2 3 .5 4 .1 N o r m . L o g

K I w t K I w t K I w t K I w t Norm. Log10 Intensity K I w t K I w t K I w t K I w t c d Glutaminolysis pathway Pentose phosphate pathway and sugar metabolism

R ib o s e - R ib u lo s e - G lu ta m in e G lu ta m a te 5 - p h o s . 5 - p h o s . R ib o s e F r u c to s e 6 .0 6 .4 4 .0 4 .0 3 .8 4 .8

5 .5 3 .5 3 .5 3 .6 4 .6 6 .2 In te n s it y In te n s it y 10 10 5 .0 3 .0 3 .0 3 .4 4 .4 6 .0 4 .5 2 .5 2 .5 3 .2 4 .2

4 .0 5 .8 2 .0 2 .0 3 .0 4 .0 N o r m . L o g N o r m . L o g K I w t K I w t K I w t K I w t K I w t K I w t

Extended Data Fig. 1 │ TCAIM KI causes an altered metabolic program in activated CD8+ T cells. Non-targeted metabolite analysis identified by GC/APCI mass spectrometry of 60 h polyclonally activated TCAIM KI or wt CD8+ T cells (n = 3). Cells had been stimulated with 5 µg/ ml plate-bound αCD3 and 2 µg/ ml soluble αCD28 at a density of 6 x 105 cells/ 200 µl culture medium. Selected metabolites of the glycolysis (a), TCA cycle (b), glutaminolysis (c), as well as pentose phosphate and sugar metabolism (d) pathway are shown. Quantitative data represents means. bioRxiv preprint doi: https://doi.org/10.1101/2021.04.20.440500; this version posted April 21, 2021. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY-NC-ND 4.0 International license.

a b Spleen Lymph nodes Colon S p le e n Lymph nodes C o lo n – hocre (KO) 8.74 25.5 1.57 13.6 73.0 0.67 2 0 2 0 8 0 1 5 1 5 6 0 C D 6 2 L

+ 1 0 1 0 4 0

62.1 82.2 4.41 5 5 2 0

6.46 19.7 1.73 14.1 69.9 2.01 0 0 0 hecre (KO) % C D 4 4

h o c reh (K e c O re ) w(K tc O re ) (K O ) h o c reh (K e c O re ) w(K tc O re ) (K O ) h o c reh (K e c O re ) w(K tc O re ) (K O ) 70.9 79.1 7.05

8.23 19.9 1.13 12.5 68.2 1.81 wtcre (KO) c C o lo n hocre (KO) hecre (KO) wtcre (KO) R2 = 0,859 R2 = 0,147 R2 = 0,348 p = 0,012 p = 0,262 p = 0,148

67.5 83.1 6.03 – CD44 8 0 8 0 8 0 CD62L 6 0 6 0 6 0 C D 6 2 L d + 4 0 4 0 4 0 S p le e n Lymph nodes C o lo n 2 0 2 0 2 0

6 0 6 0 6 0 0 0 0 % C D 4 4 5 1 0 1 5 2 0 2 5 5 1 0 1 5 2 0 2 5 5 1 0 1 5 2 0 2 5 4 0 4 0 4 0

T c e lls age (weeks) + 2 0 2 0 2 0 e S p le e n

% C D 8 Lymph nodes 0 0 0 1 2 1 2

9 9 T c e lls ) + 6 h o c reh (K e c O re ) w(K tc O re ) (K O ) h o c reh (K e c O re ) w(K tc O re ) (K O ) h o c reh (K e c O re ) w(K tc O re ) (K O ) 6 6 *1 0 ( 3 3

0 0 a b s . # C D 8

h o c reh (K e c O re ) w(K tc O re ) (K O ) h o c reh (K e c O re ) w(K tc O re ) (K O )

Extended Data Fig. 2 │TCAIM KO favors spontaneous effector differentiation of peripheral CD8+ T cells. Frequencies (a) and absolute numbers (b) of CD8+ T cells residing in the spleen, lymph node or colon of mice with homozygous (hocre), heterozygous (hecre) TCAIM KO or wild type (wtcre) gene expression measured ex vivo by flow cytometry. Representative dot plots (c) and quantification of CD62L– / CD44+ effector/ effector memory CD8 T cell frequencies (d) from TCAIM KOCd4Cre mice with the indicated genotype and tissue. (e) Correlation analysis between the mice’s age and the frequency of colon residing CD62L– / CD44+ effector/⁺ effector memory CD8 T cells from TCAIM KOCd4Cre mice with homozygous (red), heterozygous (pink) TCAIM KO or wild type (dark grey) gene expression. Squared Pearson’s correlation coefficient (R2) and p-values are⁺ shown for each one-tailed comparison. Quantitative data represents boxplots with whiskers from min to max (n = 5). bioRxiv preprint doi: https://doi.org/10.1101/2021.04.20.440500; this version posted April 21, 2021. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY-NC-ND 4.0 International license. a M y c H if1  Extended Data Fig. 3 -dependent expression 1 5 .0 1 6 KI │TCAIM + w t ( K I) of metabolism controlling genes in activated CD8

E x p r . 1 5 2 1 4 .5 KO T cells. (a) Myc and Hif1α mRNA expression of naïve and 1 4 + 1 4 .0 polyclonally activated CD8 T cells with TCAIM KI, KO or 1 3 wt gene expression (KI and wt: n = 3; KO: n = 2). RNA-seq 1 3 .5 N o r m . L o g raw count data of protein-coding genes were normalized 0 2 4 4 8 0 2 4 4 8 and log -transformed after adding a pseudocount to each time (h) 2 value. Cells had been stimulated with 5 µg/ ml plate-bound b 5 padj = 0.05 padj = 0.3 αCD3 and 2 µg/ ml soluble αCD28 at a density of 6 x 10 Suclg1 cells/ 200 µl culture medium. Quantitative data represents Sdhb Psph means ± SEM. (b) Overview of the differentially expressed Psat1 enzymes controlling the glycolysis, TCA cycle and Pkm Phgdh glutaminolysis pathway between 24 to 48 h polyclonally Pgam2 + activated TCAIM KI and wt CD8 T cells (log2 fold Pfkfb1 Pdhx change). P-values have been adjusted to either 0.05 or 0.3 Pdhb as indicated (n = 3). White areas indicate p-values Pdha1 Ogdhl exceeding the adjusted ones. Cells had been stimulated Mdh2 with 5 µg/ ml plate- µg/ ml soluble Ldhd bound αCD3 and 2 5 Ldhb αCD28 at a density of 6 x 10 cells/ 200 µl culture medium L2hgdh (a-b). Idh3a Idh1 Hk2 Gpt2 Got2 Got1 Glud1 Gls2 Gls Gapdh Eno3 Eno1 Csl Cs

24 h 48 h 24 h 48 h

Log 2 (fold change)

-0.5-1-1.5 0 0.5 bioRxiv preprint doi: https://doi.org/10.1101/2021.04.20.440500; this version posted April 21, 2021. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY-NC-ND 4.0 International license.

a H m g s c 1 H m g c r M v k Id i1 Id i2 1 5 1 5 1 0 .5 1 1 1 2 KI 1 4 1 4 1 0 .0 1 0 9 E x p r . w t ( K I) 2 9 1 3 1 3 9 .5 6 KO 8 1 2 1 2 9 .0 3 7 1 1 1 1 8 .5 0 N o r m . L o g 0 2 4 4 8 0 2 4 4 8 0 2 4 4 8 0 2 4 4 8 0 2 4 4 8

tim e ( h ) b T m 7 s f 2 E b p S c 5 d D h c r 7 D h c r 2 4 H s d 1 7 b 7

1 1 .0 1 2 .5 1 4 1 4 1 4 1 2

1 0 .5 1 2 .0 1 3 1 3 1 3 1 1 E x p r . 2 1 0 .0 1 1 .5 1 2 1 2 1 2 1 0

9 .5 1 1 .0 1 1 1 1 1 1 9

9 .0 1 0 .5 1 0 1 0 1 0 8 N o r m . L o g 0 2 4 4 8 0 2 4 4 8 0 2 4 4 8 0 2 4 4 8 0 2 4 4 8 0 2 4 4 8 time (h)

Extended Data Fig. 4 │ Reduced gene expression of mevalonate pathway and cholesterol biosynthesis controlling enzymes in TCAIM KI CD8+ T cells. mRNA expression of additional genes controlling the mevalonate pathway (a) and the cholesterol biosynthesis (b) of naïve and polyclonally activated CD8+ T cells with TCAIM KI, KO or wt gene expression (KI and wt: n = 3; KO:

n = 2). RNA-seq raw count data of protein-coding genes were normalized and log2-transformed after adding a pseudocount to each value. Cells had been stimulated with 5 µg/ ml plate-bound αCD3 and 2 µg/ ml soluble αCD28 at a density of 6 x 105 cells/ 200 µl culture medium. Data represents means ± SEM. bioRxiv preprint doi: https://doi.org/10.1101/2021.04.20.440500; this version posted April 21, 2021. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY-NC-ND 4.0 International license.

a TCAIM KI wt (KI) TCAIM KO wt (KO) 0 h

24 h

48 h

b c d K I w t KO w t K I w t KO w t ** *** 2 5 ** ** * 1 .0 4 0 *** ****** round 2 0 0 .8 3 0 1 5 0 .6 2 0 1 0 0 .4 elongated 1 0 5 0 .2

0 0 .0 # Cristae0 / M ito. Roundness factor

# mitochondria / cell K I w t K O w t 0 h 0 h 0 h 0 h 0 h 0 h 0 h 0 h 2 4 h4 8 h 2 4 h4 8 h 2 4 h4 8 h 2 4 h4 8 h 2 4 h4 8 h 2 4 h4 8 h 2 4 h4 8 h 2 4 h4 8 h

Extended Data Fig. 5 │ CD8+ T cells display an activation-dependent increase in number of mitochondria and cristae. (a) Representative transmission electron microscopy images of naïve (upper row) and 24 to 48 h (middle and bottom row, respectively) polyclonally activated CD8+ T cells with TCAIM KI, KO or respective to mouse strain wt gene expression. Scale bars: 500 nm. (b-d) Quantitative plots of transmission electron microscopy image analysis of naïve and polyclonally activated TCAIM KI, KO and respective wt CD8+ T cells. Absolute number of mitochondria per cell (b), mitochondrial roundness factor of 48 h activated CD8+ T cells only (c) and absolute number of cristae per mitochondrion (d). For each, mitochondria of 30 cells have been analyzed. Data are represented as scatter dot plots and were analyzed in accordance to Gaussian distribution of the compared sample sets by Kruskal-Wallis test with Dunn’s multiple comparisons test and subsequent post hoc one-tailed, unpaired Student’s t-test (b: KI 0 h vs. 48 h and 24 h vs. 48 h, wt 0 h vs. 24 h) or one-tailed, unpaired Mann-Whitney test (b: KO 0 h vs. 24 h; c: all). * p < 0.05, ** p < 0.01, *** p < 0.001. bioRxiv preprint doi: https://doi.org/10.1101/2021.04.20.440500; this version posted April 21, 2021. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY-NC-ND 4.0 International license.

Extended Data Table 1. Differentially expressed genes being up- or downregulated in 48 h activated TCAIM KI compared to wt CD8+ T cells for chosen annotated GO terms.

ID # up top 20 # down top 20 GO:0001816 30 Pglyrp2; P2rx7; Tlr6; F2r; Klf2; Stat1; 9 Klrk1; Optn; Il23a; Nr4a3; Lgr4; Ifng; Il6ra; Fgfr1; Rorc; Runx1; Twist1; Gbp4; Trim16; Mc1r; Il12rb2 Sh2d1b1; P2ry2; Gprc5b; Il1r2; Ceacam1; Cxcr3; Lbp; Traf3ip1 GO:0006952 56 Cxcr6; Pglyrp2; P2rx7; Ptger2; Tnfaip8l2; 18 Rab34; Klrk1; Sh2d1a; Optn; Ldlr; Tlr6; F2r; Ifit1bl1; Stat1; Gbp9; Rorc; Rab11fip5; Il23a; Stxbp1; Tbkbp1; Tnfrsf26; Gbp7; Cd86; Gbp4; Sh2d1b1; Fndc4; Fes; Ifitm3; Lgr4; Pde5a; Ifng; Gprc5b; Il1r2; Lyz2; Oas2 Unc13b; Bcl6b; Padi4 GO:0009605 65 Cxcr6; Pglyrp2; Ccr9; P2rx7; Ptger2; 24 Rab34; Klrk1; Optn; Cd80; Batf3; Scd4; Tubb3; Tnfaip8l2; Tlr6; F2r; Grm6; Ldlr; Il23a; Nr4a3; Ptprf; Tmprss6; Fndc4; Cacna2d4; Ifit1bl1; Stat1; Oprm1; Nucb2; Fes; Cdkn1a; Ifitm3; Pnpla3; Serpine2; Gbp9; Fgfr1; Rorc; Tnfrsf26; Gbp7 Pde5a; Zswim5; Ifng GO:0016126 0 9 Mvd; Fdft1; Pmvk; Fdps; Msmo1; Nsdhl; Lss; Cyp51; Sqle GO:0045595 47 Pglyrp2; Stat1; Ldlrad4; Oprm1; Nucb2; 26 Fdps; Lrp8; Gpr55; Lgals1; Acsl6; Nphp3; Timp2; Fgfr1; Rorc; Tnfrsf26; Runx1; Ldlr; Tle6; Il23a; Ptprf; Ptpn5; Fes; Lif; Igf1r; Twist1; Sgk1; P2ry2; Gprc5b; Wnt3; Spint1; Serpine2; Pde5a; Zswim5; Ceacam1; Wnt5b; Cd1d1; Car2 Epb41l5; Ifng; Trim16 GO:0048585 48 Pglyrp2; P2rx7; Tnfaip8l2; Tlr6; Stat1; 19 Tmbim1; Plaur; Optn; Adgrg3; Nphp3; Ldlrad4; Oprm1; Nucb2; Timp2; Fgfr1; Tbc1d7; Deptor; Nr4a3; Ptprf; Tmprss6; Rorc; Igf1r; Twist1; Sh2d1b1; Wnt3; Fndc4; Ptpn5; Lif; Lgr4; Serpine2; Cish; Spry1; Il1r2; Ptpre; Ceacam1; Wnt5b Svip; Dkkl1; Celf4 GO:0045597 29 Stat1; Oprm1; Timp2; Fgfr1; Runx1; Igf1r; 16 Fdps; Lrp8; Acsl6; Tle6; Il23a; Ptprf; Twist1; Sgk1; P2ry2; Gprc5b; Wnt3; Ptpn5; Fes; Lif; Spint1; Serpine2; Pde5a; Ceacam1; Wnt5b; Cd1d1; Car2; Dmbt1; Epb41l5; Ifng; Trim16; Dkkl1 Ephb2; Scin; Tgfb1i1; Shank1 GO:0001818 14 Pglyrp2; Tlr6; Klf2; Fgfr1; Twist1; Gbp4; 4 Il23a;g Lgr4; Ifn ; Mc1r Sh2d1b1; Il1r2; Ceacam1; Lbp; Traf3ip1; Gpnmb; Fn1; Cd276 Genes in bold highlight the deficiency in T cell activation and effector function of TCAIM KI CD8+ T cells.