NEDDylation is a posttranslational modification whereby the ubiquitin-like molecule NEDD8 is attached to protein substrates in a process dependent on NEDD8-activating enzyme regulatory subunit (NAE1). NEDDylation is emerging as a regulator of cancer biology, but its precise role in antitumor immunity has not been thoroughly characterized. In this study, we examine the impact of NEDDylation in CD8+ T cell–mediated antitumor responses. Analysis of publicly available single-cell RNA sequencing databases revealed that CD8+ tumor-infiltrating lymphocytes showed increased expression of NEDD8 during their differentiation into effector memory cells. In vitro activation of mouse and human CD8+ T cells drove the upregulation of the NEDDylation enzymatic pathway, resulting in an enrichment of NEDDylated proteins. In vivo tumor challenge assays demonstrated that CD8+ T cells lacking NAE1 exhibited reduced antitumor capability and a less activated phenotype with compromised differentiation into effector cells. Upregulating NEDDylation by knocking out deNEDDylase sentrin-specific protease 8 increased the in vitro cytotoxic capability of CD8+ CAR T cells. In addition, LC MS/MS proteomic analyses of NAE1-deficient CD8+ T cells and CD8+ T cells treated with the NEDDylation inhibitor MLN4924 showed a pronounced impairment in metabolic pathways, including glycolysis and oxidative phosphorylation. In this context, we validated lactate dehydrogenase A, α-enolase, and hexokinase 1, which are relevant glycolytic enzymes, as NEDD8 targets. In line with this, NEDDylation-deficient CD8+ T cells demonstrated reduced transcription, protein expression, and enzymatic activity of lactate dehydrogenase. In summary, we uncover NEDDylation as a critical regulator of CD8+ T cell–mediated antitumor immunity.

Posttranslational modifications (PTM) are central regulators of immune responses in the tumor microenvironment (1). The most extensively studied PTMs include those that trigger signaling pathways [e.g., phosphorylation cascades triggered by T-cell receptor (TCR) engagement] and those that maintain protein homeostasis, such as ubiquitination (2, 3). The latter refers to the conjugation of ubiquitin units to substrate proteins, whereas similar PTMs, including NEDDylation and SUMOylation, utilize other substrates called ubiquitin-like proteins to modify target proteins. Given that PTMs are recognized as key regulators of T-cell function (4), we aimed to study the role of NEDDylation in CD8+ T cell–mediated antitumor immunity.

NEDDylation involves the covalent attachment of the ubiquitin-like protein NEDD8 to target proteins (5). This reaction is mediated by the formation of an isopeptide bond between the carboxy-terminal glycine of NEDD8 (Gly76) and a lysine residue on the target protein (6). NEDDylation is regulated by NEDD8-activating enzyme (NAE), which is comprised of two subunits: an E1 catalytic subunit referred to as UBA3 and an E1 regulatory subunit identified as NAE1. NAE is responsible for activating NEDD8, which can then be transferred to an E2 conjugating enzyme, such as NEDD8-conjugating enzyme UBC12 (UBE2M) or NEDD8-conjugating enzyme UBE2F, which load NEDD8 into an E3 ligase complex (7). The most described targets of NEDD8 are cullin–RING ligases (CRL), whose primary role is to tag ubiquitin to specific proteins for degradation by the proteasome. This process is regulated by NEDDylation, because NEDD8 ligation to CRLs leads to a conformational change that activates the CRL complex (8, 9). Finally, deNEDDylases, such as SENP8 (also called NEDP1), are responsible for removing NEDD8 from NEDDylated substrates in a dynamic and tightly regulated cycle (10).

In addition to CRLs, NEDDylation can act on alternative protein substrates. Notably, NEDDylation can reduce the activity of the tumor suppressor P53 and promote the degradation of EGFR (11, 12). Conversely, NEDD8 attachment can increase the stability of certain proteins, such as TGF-β receptor 2 (13). Therefore, NEDDylation effects are variable and substrate-dependent (14), including protein stabilization or degradation, intracellular cell trafficking, modulation of enzymatic activity, or regulation of DNA damage response and metabolism (1518). As such, uncovering additional NEDDylation targets beyond CRLs is critical if we are to fully understand its role in health and disease.

In the context of cancer, the NEDDylation pathway has been described to be upregulated in several tumor types (1922), and it is considered a poor prognostic marker (23). The dysregulation of NEDDylation in cancer cells presents an opportunity for the development of cancer therapies (24). MLN4924 (pevonedistat) is a first-in-class small-molecule inhibitor of the E1-activating enzyme NAE, which blocks NEDDylation, disrupting protein turnover, reducing angiogenesis and proliferation, and inducing apoptotic death of tumor cells (25, 26). NAE is an adequate target for therapies that only aim to regulate NEDDylation because it exclusively interacts with NEDD8, whereas other enzymes of the NEDDylation cycle can also use ubiquitin as substrate. Currently, MLN4924 is being evaluated in clinical trials for the treatment of various cancer types (27, 28). However, the phase III clinical trial evaluating the efficacy of MLN4924 for the treatment of myelodysplastic syndromes or acute myeloid leukemia failed in 2022 to meet the primary endpoint (29). Similarly, the NAE inhibitor TAS446 also failed to complete phase I clinical trial testing against hematologic cancers (ClinicalTrials.gov identifier: NCT02978235). Given the relevance of the immune system in these pathologies (3032), elucidating the role of NEDDylation on immune cell function is of relevance, but few studies have looked into it (33, 34).

Within current knowledge, NEDDylation regulates macrophage function in the context of responses to checkpoint receptor blockade (35). Inhibition of NEDDylation suppresses the function of dendritic cells, limiting inflammatory responses (36), and reduces the activity of CD4+ T cells against blood-stage Plasmodium infection. The suppression of NEDDylation is also being explored as a method to alleviate COVID-19–associated inflammation (37) and multiple sclerosis (38). These insights reveal that modulating NEDDylation could be key in devising new treatments for immune-mediated conditions.

The precise role of NEDDylation on regulating the function of CD8+ T-cell subsets, such as effector or exhausted T cells, is not fully understood. In this study, we report that the use of a combination of in vitro, in vivo, and computational approaches showed that NEDDylation is necessary for CD8+ T-cell proliferation and differentiation into effector cells. Furthermore, abrogating NEDDylation by both genetic deletion and pharmacologic inhibition of NAE1, as well as increasing NEDDylation by the deletion of deNEDDylase SENP8 in CD8+ T cells, uncovered its critical role in regulating CD8+ T-cell fate, metabolism, and antitumor function.

Mice

C57/BL6 female mice were purchased from Charles River Laboratories, whereas dLCK-Cre mice (C57/BL6 background) were purchased from The Jackson Laboratory (strain #:012837). NAE1fl/fl mice (C57/BL6 background) were kindly provided by Dr. Ashwin Whoodoo (University of Santiago de Compostela, Spain; ref. 39). dLCK-Cre and NAE1fl/fl mice were crossed to obtain NAE1 knockout (KO) mice. Cre negative littermates were used as controls. OT-I mice were obtained from The Jackson Laboratory (strain #:003831, C57/BL6 background) and were crossed with NAE1-KO mice to generate NAE1-KO-OT-I mice, which were used for antigen-specific assays. Animal procedures were performed following the ethical guidelines established by the Biosafety and Welfare Committee at CIC bioGUNE and the recommendations from Association for Assessment and Accreditation of Laboratory Animal Care International. These experiments have been approved by the Committee on Bioethics and Animal Welfare (CBBA) and by the government of Bizkaia (Spain) under license numbers “P-CBG-CBBA-0319” and “P-CBG-CBBA-521.”

In vitro assays

Culture of mouse and human primary CD8+ T cells

For the isolation of mouse CD8+ T cells, individual spleens from NAE1-KO and NAE1-KO-OT-I mice were homogenized using a syringe plunger in 20 mL of RPMI 1640 Glutamax medium (Gibco, cat. #: 11875093), and a 70 µm strainer was used for further dissociation. The single-cell suspension was centrifuged (5 minutes at 600 g, 4°C). Red blood cells were lysed by resuspending the pellet in 3 mL of ammonium-chloride-potassium (ACK) lysis buffer (Thermo Fisher Scientific, cat. #: A1049201). After 3 minutes, 10 mL of RPMI 1640 Glutamax medium were added, and the cell suspension was centrifuged (5 minutes at 600 g, at 4°C). Splenocytes were resuspended in RoboSep buffer (Stemcell Technologies, cat. #: 20104) and counted. CD8+ T cells were magnetically isolated using EasySep Mouse CD8+ T cell isolation Kit (Stemcell Technologies, cat. #: 19853A). Isolated CD8+ T cells were activated using plate-bound anti-CD3 (5 μg/mL, Thermo Fisher Scientific, cat. #: 16-0031-85) and soluble anti-CD28 (1 μg/mL, Thermo Fisher Scientific, cat. #: 16-0281-85). Cell culture medium consisted of 1640 RPMI Glutamax medium supplemented with β-mercaptoethanol (50 μmol/L, Thermo Fisher Scientific, cat. #: 31350010) and HEPES (25 mmol/L, Thermo Fisher Scientific, cat. #: 15630056). CD8+ T-cell culture density was 1.5 × 106 cells per well (24-well plates) or 0.2 × 106 per well (96-well plates) at a concentration of 106 cells/mL. DMSO was added to control wells. After 3 days of culture, CD8+ T cells were washed three times and reseeded with cell culture medium supplemented with IL2 (20 ng/mL, BioLegend cat. #: 575404). When indicated, NEDDylation inhibitor MLN4924 (MedChemExpress, cat. #: HY-70062) was added at 0.2 μmol/L or 0.4 μmol/L concentrations. When mentioned, galactose (10 mmol/L, Sigma-Aldrich, cat. #: 5388-100G) was added to no-glucose RPMI medium (Gibco cat. #: 11879020). For the analysis of thymocytes, thymi were harvested and homogenized with a syringe plunger in RPMI 1640 Glutamax medium. For the analysis of inguinal lymph nodes, these were homogenized with a syringe plunger in RPMI 1640 Glutamax medium. Cells were directly stained and analyzed by flow cytometry.

For the isolation of human CD8+ T cells, buffy coats were provided by the Basque Biobank (www.biobancovasco.org) and processed following standard operation procedures with appropriate approval of ethical and scientific committees (code CEIC E19-75). Human peripheral blood mononuclear cells were isolated from buffy coats using Ficoll-Paque Plus (Cytiva, cat. #: 17144003) density centrifugation at 1,200 g for 10 minutes at room temperature (RT) using SepMat-50 tubes (Stemcell Technologies, cat. #: 85450). CD8+ T cells were magnetically isolated using EasySep Human CD8+ T Cell Isolation Kit (Stemcell Technologies, cat. #: 17953RF) in an automated RoboSep cell separator (Stemcell Technologies) according to the manufacturer’s indications. Human cD8+ T cells were stimulated on culture plates coated with anti-CD3 (5 μg/mL, BD Biosciences, cat. #: 567118) and soluble anti-CD28 (1 μg/mL, Thermo Fisher Scientific, cat. #: 16-0289-81) in OpTmizer CTS medium (Thermo Fisher Scientific, cat. #: A1048501) supplemented with CTS Immune Cell Serum Replacement (Thermo Fisher Scientific, cat. #: A2596101). Cells were also stimulated with phorbol 12-myristate-13-acetate (PMA, 200 ng/mL, Sigma-Aldrich, cat. #: P8139) and ionomycin (200 ng/mL, Sigma-Aldrich, cat. #: I0634). Cells were seeded at a density of 106 cells/mL of medium. DMSO was added in control wells. After 3 days, CD8+ T cells were washed three times and refreshed with new media. When indicated, MLN4924 was added at 0.2 or 0.4 μmol/L concentrations.

Cell lines

LLC (cat. #: CRL-1642; 2019) and B16 cells (cat. #: CRL-6475; 2019) were purchased from the ATCC, cultured in DMEM (Gibco cat. #: 41966029), and passaged every 2 days. The B16-OVA cell line was a gift from Dr. Ignacio Melero (Centro de Investigación Médica Aplicada, University of Navarra, Spain) and passaged every 2 days. The rest of the ovalbumin (OVA) and GFP luciferase-expressing cell lines were generated through lentiviral infection. HEK 293 cells (Takara Bio Inc., cat. #: 632180; 2020) were cultured in DMEM (Gibco, cat. #: 41966029) and passaged every 2 days. YUMMER 1.7-H2B-GFP5 cells were purchased from Sigma-Aldrich (cat. #: SCC245; 2024), cultured in DMEM-F12 (Gibco cat. #: 11320033) plus nonessential amino acids (Gibco, cat. #: 11140035), and passaged every 3 days. RAMOS cells were generously provided by Dr. Francisco Borrego (Biobizkaia Health Research Institute, Spain), cultured in RPMI 1640 Glutamax medium, and passaged every 2 days. Jurkat (clone E6-1) cells (cat. #: TIB-152, 2019) were purchased from the ATCC, cultured in RPMI 1640 Glutamax medium, and passaged every 2 days. The JE6.1 triple parameter reporter (Jurkat TPR) cell line was generously provided by Dr. Peter Steinberger (Medical University of Vienna, Austria), cultured in RPMI 1640 Glutamax medium, and passaged every 2 days. Jurkat TPR cells are genetically modified to encode fluorescent proteins under the control of response elements for transcription factors like NF-κB (eCFP proteins), NFAT (eGFP proteins), or AP-1 (mCherry proteins). All cell lines were cultured in media supplemented with 10% FBS (Thermo Fisher Scientific, cat. #: 10270106) and 1% penicillin–streptomycin (Thermo Fisher Scientific, cat. #: 15140122) and tested for Mycoplasma using MycoAlert Mycoplasma Detection Kit (Lonza, cat. #: LT27-221) following the manufacturer’s instructions. Microscopic examination to check cell shape, size, and growth patterns was performed daily. Cells were maintained in an incubator at 37°C in a 5% CO2 atmosphere and cultured for a maximum of 10 passages.

Plasmid design and cloning

Plasmid encoding OVA was a gift from Maria Castro (University of Michigan, United States, Addgene plasmid # 25097; http://n2t.net/addgene:25097; RRID: Addgene_25097), and the OVA region was cloned into MSCV-pLV lentiviral plasmid by GenScript Biotech (resulting plasmid: cOVA-puro-MSCV-pLV).

For CAR T-cell generation, we used CAR constructs which included the CD8α signaling peptide (UniProt entry: P01732), the anti-CD19 scFv derived from the FMC63 mAb (40), the CD8α hinge/transmembrane domain (UniProt entry: P01732), the 4-1BB costimulatory domain (UniProt entry: Q07011), the CD3ζ intracellular domain (UniProt entry: P20963-3), and a truncated form of the nerve growth factor receptor (NGFRt) as a selection marker (UnipPot entry: P08138). NGFRt was expressed separately to the CAR via a T2A self-cleaving peptide. This construct was cloned into the third-generation lentiviral vector pCCL (41), under the control of the EF1a constitutive promoter and synthesized by Genscript, resulting in the plasmid FMC63-BBz-NGFRt pCCL (item no. U2706FF040-10). For SENP8-KO plasmid production, plasmid encoding “lentiCRISPR v2” was a gift from Feng Zhang (Broad Institute, United States; Addgene plasmid cat. #: 52961). The guide RNAs against human AAVS1 (safe harbor gene used as control: GGG​GCC​ACT​AGG​GAC​AGG​AT) and SENP8 (AAC​TCA​GTT​CAC​GCA​AAG​C; ref. 42) were cloned into this vector by GenScript Biotech, resulting in plasmids “hAAVS1 KO_(ctrl)_LentiCRISPRv2” and “NEDP1_2_GFP_PURO_Lenticrisprv2,” respectively. GFP as a visualization marker and the puromycin resistance gene as a selection marker were also included in the plasmid.

The heat shock method was used to incorporate plasmids into 15 μL of One Shot TOP10 (Thermo Fisher Scientific, cat. #: C404003) bacteria. Briefly, 0.3 μg of plasmid was added to 15 μL of bacteria and left 20 minutes on ice. After this, cells were heated to 42°C for 1.5 minutes and then incubated on ice for an additional 5 minutes. Bacteria were grown overnight for 16 hours in agar plates with ampicillin (100 μg/mL, Sigma-Aldrich, cat. #: A9518) as the selection antibiotic. The selected colony was used to inoculate a starter culture, which was grown overnight for 16 hours. Subsequently, the starter culture was used to inoculate a larger culture for maxiprep during additional 16 hours. Then, DNA was isolated using HiPure PureLink Maxiprep Isolation Kit (Thermo Fisher Scientific, cat. #: K210007).

Lentivirus generation and T-cell transductions

To generate lentiviral particles, 5 × 106 HEK 293T cells were seeded in 10-cm plates and transfected on the next day using jetPEI Kit (Polyplus Transfection, cat. #: 101-10N) using packaging plasmids psPAX2 (4 μg, a gift from Didier Trono, École Polytechnique Fédérale de Lausanne, Switzerland, Addgene plasmid cat. #: 12260), and VSV-G (1.5 μg, a gift from Tannishtha Reya, Columbia University, United States, Addgene plasmid cat. #: 14888). For the rest of the plasmids, 5 μg per transfection plate was used. Viral particles were harvested from the supernatant 48 hours after transfection, centrifuged, and filtered through a 0.45-μm filter (VWR cat. #: 514-0063).

Virus used for T-cell transductions was concentrated using a Lenti-X concentrator (Takara Bio, cat. #: 631232) and titrated on Jurkat cells. Briefly, 5 × 104 Jurkat cells were seeded in 96-well plates with 8 μg/mL of polybrene (Sigma-Aldrich, cat. #: TR-1003). A serial dilution of the used virus was added to the wells. Each dilution was assayed in duplicate. After 2 days, the efficiency of infection was analyzed by flow cytometry. Only the range of 5% to 20% of infection was used to quantify the amount of active viruses per μL.

Generation of stable cell lines through lentiviral infection

For the generation of stable cell lines, LLC, B16, and B16-OVA cells were seeded in a 6-well plate, and 24 hours later, nonconcentrated lentiviral particles were added in the presence of polybrene at 5 µg/mL for LLC and 10 µg/mL for B16 and B16-OVA. After 2 days, the expression of the desired antigen was checked by flow cytometry. For luciferase-expressing cells (B16-GFP-luciferase and B16-GFP-luciferase-OVA), transduced cells were sorted using a BD FACSAria Fusion sorter (BD Biosciences, purity >95%). For OVA-expressing LLC cells (LLC-OVA), puromycin selection was performed (2 µg/mL, Thermo Fisher Scientific, cat. #: A1113803).

T-cell lentiviral transduction

Human CD8+ T cells isolated from healthy blood donors were activated with anti-CD3 and anti-CD28 Dynabeads (1:1 ratio, Thermo Fisher Scientific, cat. #: 11132D) at 106 cells/mL with the mentioned human T-cell medium supplemented with IL2 (100 IU/mL, Miltenyi Biotech, cat. #: 130-097-746). On day 1 of activation, 0.5 × 106 cells were seeded in 24-well plates and transduced with concentrated lentivirus. First, hAAVS1-KO (control) or SENP8-KO viral infection was performed at a multiplicity of infection of 6. The medium mentioned above contained polybrene at 8 μg/mL. Spinoculation was performed (2,000 RPM for 1 hour at 32°C). At day 3 of activation, transduced cells were selected with puromycin (1 μg/mL) for 3 days. Dynabeads were removed after 6 days. Every 2 days, IL2 was replenished with fresh media and cells were left at 0.5 × 106 cells/mL. After 15 days of culture, cells were reactivated with anti-CD3 and anti-CD28 Dynabeads following the same protocol as stated before. After 1 day of activation, CD8+ T cells were infected with anti-CD19 CAR virus at a multiplicity of infection of 3. Cells were expanded for 15 days, after which they were purified using the NGFR positive selection kit (Stemcell Technologies, cat. #: 17849), obtaining purities greater than 80%. After 2 days, SENP8-KO anti-CD19 CAR T cells were used for the cytotoxicity assay.

In vitro T-cell cytotoxicity assay

For the mouse OT-I–based cytotoxicity assay, OT-I CD8+ T cells from control or NAE1-KO OT-I mice were magnetically isolated from the spleen and were activated in vitro for 3 days. At day 2, 2 × 104 B16-luciferase-OVA cells were seeded in a white 96-well plate (PerkinElmer, cat. #: 6005680). At day 3, activated OT-I CD8+ T cells were washed and added to the culture at the indicated effector cell:T cell ratios for 24 hours.

For the human SENP8-KO anti-CD19 CAR T cell cytotoxicity assay, 2.5 × 104 RAMOS-luciferase cells were seeded in the same plates as the previous experiments. On the same day, hAAVS1 or SENP8-KO anti-CD19 CAR T cells were added to each well at the indicated effector cell:T cell ratios for 24 hours.

In both assays, at endpoint 50 μL of D-Luciferin (3 mg/mL, PerkinElmer, cat. #: 122799) were added to each well, and the luminescence signal corresponding to alive cells was measured using a multimode plate reader (PerkinElmer Victor Nivo).

In vitro T-cell exhaustion

For the in vitro T-cell exhaustion model, we followed the protocol previously established by Scharping and colleagues (43). Briefly, splenic CD8+ T cells were isolated from control and NAE1-KO mice. Cells were activated at 1 × 106 cells/mL using anti-CD3 and anti-CD28 Dynabeads in a 1:1 ratio, in RPMI 1640 Glutamax medium supplemented with 10% FBS, 1% penicillin–streptomycin, 25 ng/mL IL2, and 10 ng/mL IL12 (Miltenyi Biotech, cat. #: 130-096-708). After 24 hours (day 1), beads were magnetically removed, and the cells were divided into two conditions: acute activation (without Dynabeads) and chronic activation (by readding Dynabeads). Fresh media supplemented with 25 ng/mL of IL2 was added. On days 3 and 5, the cells were split in half with fresh IL2 supplemented media. Finally, on day 6, the cells were analyzed.

Measurement of IFNγ production

IFNγ levels in cell culture supernatants were quantified after all the stimulations of the T-cell exhaustion protocol using a sandwich ELISA kit according to the manufacturer’s protocol (R&D Systems, cat. #: DY485).

Lactate dehydrogenase activity assay

The activity of lactate dehydrogenase (LDH) in cell culture supernatants was quantified on day 3 of activation according to the manufacturer’s colorimetric protocol (Sigma-Aldrich, cat. #: MAK066).

Flow cytometry

Single-cell suspensions were seeded in V-shaped 96-well plates (Thermo Fisher Scientific, cat. #: 249570). If cells were to be fixed, they were first stained with fixable LIVE/DEAD viability dyes for 30 minutes at 4°C protected from light. Then, cells were incubated with TruStain FcX blocker for 10 minutes at RT protected from light. After this, cells were stained with fluorochrome-conjugated antibodies in staining buffer (Thermo Fisher Scientific, cat. #: 00-4222-26) for 30 minutes at 4°C in the dark. After staining, if cells were to be fixed, they were resuspended in paraformaldehyde at 0.5% and acquired no more than 5 days later in a BD FACSymphony A3 flow cytometer. If cells were not fixed, 4′,6-diamidino-2-phenylindole (DAPI,1:10,000 dilution, Thermo Fisher Scientific, cat. #: D1306) was added as a viability marker. Analysis was performed using FlowJo v10 or Cytobank Premium software. All centrifugation steps were performed at 450 g for 5 minutes at 4°C. The antibodies used are listed in Supplementary Table S1. Specific staining procedures are detailed in the corresponding sections of “Materials and Methods.”

Cell proliferation and viability assay

Absolute cell numbers of mouse or human CD8+ T cells were obtained by using CountBright absolute counting beads (Thermo Fisher Scientific, cat. #: C36995) after acquisition in a BD FACSymphony A3 flow cytometer.

For the carboxyfluorescein diacetate succinimidyl ester (CFSE) dilution assay of mouse CD8+ T cells, 106 CD8+ T cells were resuspended in RPMI medium supplemented with FBS and P/S and were incubated with 5 μmol/L CFSE (Thermo Fisher Scientific, cat. #: C34554) for 5 minutes at RT in the dark. Cells were washed three times and 3.5 × 105 CD8+ T cells were activated with plate-bound anti-CD3 and soluble anti-CD28 in round-shaped 96-well plates. At day 3, cells were acquired in a BD FACSymphony A3 flow cytometer.

For the CFSE dilution assay of human CD8+ T cells, 106 CD8+ T cells were resuspended in PBS and incubated with 5 μmol/L CFSE for 7 minutes at 37°C in the dark. Five mL of ice-cold RPMI were added to the cells, and the cells were incubated for 5 minutes and washed twice. Then, 1.75 × 105 CD8+ T cells were activated with plate-bound anti-CD3 and soluble anti-CD28 in a round-shaped 96-well plate and acquired in a BD FACSymphony A3 flow cytometer at day 4 of activation.

For the annexin V viability assay, activated mouse CD8+ T cells were harvested after 3 days of culture and stained with 3 µL of annexin V-BV421 in 100 µL of binding buffer (BD Biosciences, cat. #: 556454). After incubating for 5 minutes in the dark, 2 µL of 7-aminoactinomycin D (7AAD) in 100 µL of binding buffer were added to samples, which were kept in ice until acquisition in a BD FACSymphony A3 flow cytometer. For human CD8+ T cells, DAPI-negative cells were considered alive cells. Data analysis was performed with FlowJo v10 or Cytobank software programs.

Western blot

Total cell lysates were collected using lysis buffer (50 mmol/L Tris, pH 8.0, 150 mmol/L NaCl, 1% Triton X-100, and 2 mmol/L EDTA) supplemented with protease inhibitors (cOmplete EDTA-free tablets, EASYPack, Roche, cat. #: 04693132001), phosphatase inhibitors (PhosSTOP EASYpack, Roche, cat. #: 04906837001), phenylmethylsulfonyl fluoride (Cell Signaling Technologies, cat. #: 8553), and deNEDDylase inhibitor 2-iodoacetamide (2 mg/mL, Sigma-Aldrich, cat. # 8047440025). Protein quantification was performed using Pierce BCA Protein Assay Kit (Thermo Fisher Scientific, cat. #: 23227). Samples were mixed with LDS sample buffer (Thermo Fisher Scientific, cat. #: NP0007) containing ditiothreitol (DTT; Melford, cat. #: MB1015), heated for 10 minutes at 95°C, and resolved in a 7.5% or 4% to 15% Mini-PROTEAN TGX precast protein gels (Bio-Rad, cat. #: 4561023 and cat. #: 4561085, respectively) with 1X Tris/glycine/SDS electrophoresis buffer (Bio-Rad, cat. #: 1704156) and PageRuler Plus Prestained Protein Ladder (Thermo Fisher Scientific, cat. #: 26619). Protein transfer to 0.2 µm PVDF membranes (Bio-Rad, cat. #: 1704156) was carried out in a Trans-Blot Turbo transfer system (Bio-Rad). Membranes were blocked for 5 minutes in EveryBlot Blocking Buffer (Bio-Rad, cat. #:12010020), washed in T-PBS (0.5% Tween-20, Sigma-Aldrich, cat. #: P2287), and incubated overnight with primary antibodies (Supplementary Table S1). The following day, membranes were washed and incubated with the corresponding secondary HRP-conjugated antibodies for 1 hour (Supplementary Table S1). Chemiluminescence detection was performed using Clarity Max Western ECL Substrate (Bio-Rad, cat. #: 170506) in an iBright CL1500 system (Thermo Fisher Scientific). Densitometric quantifications were performed using ImageJ–Fiji software. When a membrane was reprobed with an additional antibody, it was incubated overnight and developed immediately after probing with the specific antibody, following the same methodology as previously described.

RT-qPCR

Total RNA was isolated from CD8+ T cells using NucleoSpin RNA Kit (Macherey-Nagel, cat. #: 740955.250). A total of 325 ng of RNA were used to synthesize the cDNA with the M-MLV reverse transcriptase (Thermo Fisher Scientific, cat. #: 28025-013) and random oligos (Thermo Fisher Scientific, cat. #: 58875). qPCR reactions were performed in triplicate in a ViiA 7 Real-Time PCR system (Thermo Fisher Scientific) from 1 μL of cDNA using PerfeCTa SYBR Green SuperMix Reagent (Quantabio cat. #: 95056-500) and gene-specific primers (Supplementary Table S1). The 2−ΔΔCt method was used to calculate the relative gene expression considering the 18S or Rplp0 genes as housekeeping genes for human and mouse species, respectively.

Immunofluorescence

For immunofluorescence, subcutaneous LLC tumors were harvested and flash-frozen in optimal-cutting-temperature medium at −80°C. Sections of 8 μm were made with a Leica CM 1860 UV cryostat and fixed with 4% paraformaldehyde (Thermo Fisher Scientific, cat. #: J61899.AP) for 10 minutes. Slides were washed twice with PBS and permeabilized with 0.1% Triton X-100 (Sigma-Aldrich, cat. #: T9284) for 10 minutes. After washing, tissue slides were incubated with 2.5% normal goat blocking serum (Vector Laboratories, cat. #: S-1012-50) for 30 minutes at RT to reduce nonspecific staining. Serum was washed, and samples were incubated at 4°C overnight with primary antibodies against NEDD8, CD3, or Ki-67 (Supplementary Table S1). The next day, slides were washed five times with PBS and stained for 45 minutes with secondary antibodies at RT (goat anti-rat Alexa Fluor 488 and goat anti-rabbit Alexa Fluor 647; Supplementary Table S1). Slides were washed again, stained with DAPI for 10 minutes at RT, and mounted with Prolong Gold Antifade reagent (Thermo Fisher Scientific, cat. #: P36930). Images were obtained using a confocal microscope (Leica SP8 Lightning, HC PL APO objective lens with 100X magnification) and analyzed using LasX software.

Seahorse XF assay

Human CD8+ T cells were activated and cultured with MLN4924 at the indicated doses as stated in the Cell Culture section. Mouse control and NAE1-KO mouse CD8+ T cells were activated and cultured as stated in the “Culture of mouse and human primary CD8+ T cells” section of this article. At day 3, cells were washed and resuspended in Seahorse XF RPMI medium (Agilent, cat. #: 103576-100) supplemented with 10 mmol/L glucose (Agilent, cat. #: 103577-100), 1 mmol/L sodium pyruvate (Agilent, cat. #: 103578-100), 2 mmol/L L-glutamine (Gibco, cat. # 25030-024), and MLN4924 at the indicated doses. Seahorse XF24 Cell Culture Microplates were coated with poly-L-lysine (50 μg/mL, Sigma-Aldrich, cat. #: P8920) for 2 hours. A total of 105 or 2 × 105 CD8+ T cells were seeded per well for the Seahorse XF Glycolytic Rate Assay (Agilent, cat. #: 103344-100) and Seahorse XF Cell Mito Stress Test, respectively. After centrifuging at 400 g for 5 minutes, cells were left 15 minutes resting at 37°C in a non-CO2 incubator. A measure of 400 μL of supplemented medium were very carefully added to each well. For the Seahorse XF Glycolytic Rate Assay, rotenone and antimycin A (0.5 μmol/L) and 2-deoxy-D-glucose (2-DG, 500 mmol/L) were used. Basal and compensatory glycolysis calculations were made through the Seahorse Analytics software. For the Seahorse XF Cell Mito Stress Test, the following drugs were used: oligomycin (1.5 μmol/L, Sigma-Aldrich, cat. #: O4876), BAM15 (2.5 μmol/L, MedChemExpress, cat. #: HY-110284), and rotenone (0.5 μmol/L, Merck, cat. #: 557368) and antimycin A (0.5 μmol/L, Sigma-Aldrich, cat. #: A8674). Basal, maximal respiration and spare respiratory capacity were calculated through the Seahorse Analytics software. Both assays were run in the Seahorse XF24 Extracellular Flux Analyzer (Agilent Technologies).

Immunoprecipitation

For each immunoprecipitation (IP) reaction, 1.5 mg of protein lysate and 10 μg of antibody were used (anti-NEDD8 or IgG Isotype control, listed in Supplementary Table S1). Human CD8+ T cells isolated from buffy coats from healthy donors were activated in vitro. Various healthy donors were used in each IP assay, and no donors were repeated in each assay. Activated CD8+ T cells were lysed according to the protocol stated in the “Western blot” section of this article. Antibodies were coupled to Protein G magnetic Sepharose beads (50 μL, Cytiva, cat. #: 28951379) for 1 hour at 4°C. After this, antibodies were cross-linked to the beads with BS3 compound (Thermo Fisher Scientific, cat. #: 21580) according to the manufacturer’s protocol. Protein lysates were precleaned with Protein G magnetic Sepharose beads coupled to 5 μg of IgG antibody (not cross-linked) during 1 hour at 4°C. Following this, lysates were incubated with the cross-linked bead–antibody complex overnight at 4°C. Subsequently, beads were washed six times, eluted in 30 μL of LDS sample buffer (Thermo Fisher Scientific, cat. #: NP0007) containing DTT (Melford, cat. #: MB1015), and boiled at 95°C for 10 minutes. Eluates were analyzed by Western blotting or by mass spectrometry (MS). In the case of MS analysis, lysis buffer did not include 2-Iodoacetamide.

Proteomic analyses

IP: SDS-PAGE followed by in-gel tryptic digestion

Immunoprecipitated protein samples were boiled in a buffer containing 50 mmol/L Tris, pH 6.8, 5% glycerol, 1.67% β-mercaptoethanol, 1.67% SDS, and 0.0062% bromophenol blue for 5 minutes and resolved in 4% to 12% gradient acrylamide gels (NuPAGE Bis-Tris gels 1.0 × 12 well, Thermo Fisher Scientific) using an XCell SureLock Mini-Cell Electrophoresis System (Thermo Fisher Scientific). A constant voltage of 150 V was applied for 45 minutes. Gels were fixed in a solution containing 10% acetic acid and 40% ethanol for 30 minutes and stained overnight in SYPRO Ruby (Bio-Rad). Gels were then washed in a solution containing 10% ethanol and 7% acetic acid for 30 minutes, and the image was acquired using a Chemidoc gel imager system (Bio-Rad). Each lane was subjected to tryptic digestion. Gel bands were washed in Milli-Q water. Reduction and alkylation were performed using 10 mmol/L DTT in 50 mmol/L ammonium bicarbonate at 56°C for 20 minutes, followed by iodoacetamide (50 mmol/L iodoacetamide in 50 mmol/L ammonium bicarbonate) for another 20 minutes in the dark. Gel pieces were dried and incubated with trypsin (12.5 μg/mL in 50 mmol/L ammonium bicarbonate) for 20 minutes on ice. After rehydration, the trypsin supernatant was discarded. Gel pieces were hydrated with 50 mmol/L ammonium bicarbonate and incubated overnight at 37°C. After digestion, acidic peptides were cleaned with TFA 0.1% and dried out in a RVC2 25 speedvac concentrator (Christ). Peptides were resuspended in 10 μL 0.1% FA and sonicated for 5 minutes prior to LC-MS/MS analysis.

General proteomic analysis: sorting for alive cells followed by in-solution tryptic digestion

Activated mouse CD8+ T cells were sorted to isolate alive cells using a BD FACS Aria fusion sorter. Total protein from activated alive mouse CD8+ T cells was extracted by incubating the extracts in a buffer containing 7 mol/L urea, 2 mol/L thiourea, and 4% CHAPS. Samples were incubated in this buffer for 30 minutes at RT with agitation and digested following the FASP protocol described by Wisniewski and colleagues (44) in 2009 with minor modifications. Trypsin was added in 50 mmol/L ammonium bicarbonate to a trypsin:protein ratio of 1:10, and the mixture was incubated for overnight at 37°C. Peptides were dried out in an RVC2 25 speedvac concentrator (Christ) and resuspended in 0.1% FA. Peptides were desalted and resuspended in 0.1% FA using C18 stage tips (Millipore) prior to acquisition.

MS analysis

Samples were analyzed in a timsTOF Pro with PASEF (Bruker Daltonics) coupled online to an Evosep ONE liquid chromatograph (Evosep). timsTOF was operated in data-dependent acquisition mode, and the data were acquired using the default data-dependent acquisition PASEF-standard_1.1sec_cycletime method provided by Bruker. Data were acquired in positive mode. Scan range was established between 100 and 1700 m/z, and 1/K0 was between 0.60 and 1.60 V·s/cm3, with a total ramp and accumulation time of 100 ms (100% duty cycle). Collison energies applied were 20 and 59 eV for the lower and upper limits of the mobility range, respectively. Charge states between 0 and 5 were specifically selected for analysis. Total PASEF ramps were 10, to a final cycle time of 1.16 seconds.

A total of 200 ng were directly loaded onto the Evosep ONE and resolved using the default 30 samples-per-day protocol (44 minutes gradient flow and 48 minutes total cycle time). The analytic column was equilibrated at 1,500 nL/minute, and the gradient flow was 500 nL/minute, increased to 1,500 nL/minute for washing. Solvents A and B were water with 0.1% formic acid (FA) and acetonitrile with 0.1% FA, respectively.

Protein identification and quantification was carried out using PEAKS X software (Bioinformatics solutions) or MaxQuant. Searches were carried out against a database consisting of Mus musuculus entries from UniProt Swiss-Prot. Precursor and fragment tolerances of 20 ppm and 0.05 Da were considered for the searches, respectively. Only proteins identified with at least one peptide at FDR < 1% were considered for further analysis. Total protein extracts were compared by means of a Student t test. Protein abundances were compared between control and immunoprecipitated samples. Immunoprecipitated proteins with at least a twofold difference were kept for further analyses. For Gene Ontology Biological Processes selection, the list of immunoprecipitated proteins was examined using the website software (https://geneontology.org/), and the most enriched processes with at least 10 proteins associated with each process were selected.

Functional analysis

Ingenuity pathway analysis (IPA, QIAGEN) was used for a characterization of the molecular events lying behind the differential protein patterns under analysis. The calculated P values for the different analyses performed determine the probability that the association between proteins in the dataset and a given process, pathway, or upstream regulator is explained by chance alone, based on a Fisher exact test (P value < 0.05 being considered significant). The activation z-score represents the bias in gene regulation that predicts whether the upstream regulator exists in an activated (positive values) or inactivated (negative values) state, based on the knowledge of the relation between the effectors and their target molecules. For published figures based on these data, the list of proteins associated with each pathway was provided by IPA software.

In vivo assays

In vivo tumor growth

For the immunophenotyping of tumor infiltrates, 5 × 105 LLC cells or 106 YUMMER 1.7-GFP cells were subcutaneously injected in 100 μL of PBS or saline solution, respectively, in the right flank of 6- to 10-week-old control and NAE1-KO mice. Tumor growth was monitored every other day using a digital caliper. All mice were sacrificed when the tumor diameter of one mouse reached 15 mm, and single-cell suspensions from spleens, tumors, and tumor-draining lymph nodes were prepared for flow cytometry staining and analysis. Spleens and lymph nodes were dissociated using a syringe, whereas mouse tumor samples were finely minced with a scalpel blade in a Petri dish. Tumors were incubated for 30 minutes at 37°C with type I collagenase (0.5 mg/mL, Sigma-Aldrich, cat. #: C0130) in DMEM. Cell suspensions were filtered through a 70-μm cell strainer. Red blood cell lysis of spleen and tumor samples was performed using ACK buffer (Thermo Fisher Scientific, cat. #: A1049201) for 3 minutes at RT. After counting, 2 × 106 cells from each tumor and spleen were stained with a LIVE/DEAD fixable blue or green dead cell stain kit for 30 minutes at 4°C protected from light. Then, cells were washed in PBS and incubated with TruStain FcX blocker for 10 minutes at RT, protected from light. After this, cells were stained with fluorochrome-conjugated antibodies in staining buffer for 30 minutes at 4°C in the dark. The antibodies used are listed in Supplementary Table S1. After staining, cells were washed and fixed with paraformaldehyde 0.5% and acquired using a BD FACSymphony A3 flow cytometer. Data analysis was performed with FlowJo v10 or Cytobank Premium software programs. Absolute cell numbers were calculated using CountBright absolute counting beads. All centrifugation steps were performed at 450 g for 5 minutes at 4°C.

For the measurement of transcription factors, tumor cell suspensions were stained following the manufacturer’s protocol of the intracellular transcription factor staining kit (Thermo Fisher Scientific, cat. #: 00-5523-00). For cytokine and granzyme B production measurement, prior to the mentioned intracellular staining, tumor cell suspensions were stimulated with PMA and ionomycin, plus protein transport inhibitors (Thermo Fisher Scientific, cat. #:00-4975-93), or as negative control, only protein transport inhibitors (Thermo Fisher Scientific, cat. #: 00-4980-93) for 4 hours.

For antigen-specific experiments, 5 × 105 LLC-OVA cells were subcutaneously injected in the right flank of 6- to 8-weeks-old C57/BL6 females. Nine days after tumor injection, OT-I CD8+ T cells from the spleens of donor control or NAE1-KO mice were sorted to enrich viable cells in a BD FACS Aria Fusion sorter. A measure of 9 × 105 OT-I CD8+ T cells were injected intravenously in the tail vein of the recipient mice. Donor OT-I CD8+ T cells had been previously activated in vitro for 3 days with anti-CD3 and anti-CD28. Mice bearing tumors bigger than 35 mm3 at the day of injection were included in the study and randomized according to their size in the different experimental conditions. Mice were monitored for tumor growth every other day using a digital caliper. Tumor volumes were calculated using the following formula: V (mm3) = (π × L × W2)/6 (L is tumor length and W is tumor width). Mice were sacrificed when tumor length reached 15 mm.

In silico single-cell RNA sequencing analysis

We first examined NEDD8 expression in tumor-associated immune and nonimmune cell populations using a publicly available single-cell lung cancer atlas, which includes 185 human primary non–small cell lung cancer samples. To complement this analysis, we assessed three additional datasets: 58 lung adenocarcinoma samples from 44 patients (GSE131907), 31 melanoma tumors (GSE115978), and 12 hepatocellular carcinoma samples (GSE125449). In these datasets, we specifically investigated NEDD8 expression in naïve and effector-like CD8+ T cells. Briefly, all datasets underwent quality control filtering based on the number of detected features and the percentage of mitochondrial gene expression. Data were then normalized, and pure CD8+ T cells were identified using scGate. Naïve and effector-like CD8+ T cells were subsequently annotated based on the human CD8+ tumor-infiltrating lymphocyte (TIL) atlas using ProjecTILs. All computational analyses were performed in R (version 4.4.1) using the following packages: Seurat (v5.1.0), UCell (v2.7.7), ProjecTILs (v3.3.0), scGate (v1.6.2), Biostrings (v2.70.3), dplyr (v1.1.4), stringr (v1.5.1), patchwork (v1.2.0), tidyverse (v2.0.0), MetBrewer (v0.2.0), RColorBrewer (v1.1-3), Matrix (v1.7-0), and biomaRt (v2.58.2).

Data availability

The input data processed in the single-cell RNA sequencing (scRNA-seq) reanalysis comprised three datasets acquired from the NCBI Gene Expression Omnibus: GSE131907 (lung cancer), GSE115978 (melanoma), and GSE125449 (hepatocarcinoma). Additionally, the integrated lung cancer atlas by Prazanowska and Lim (45) was downloaded from https://figshare.com/articles/dataset/NSCLC_Final_dataset/22114682?backTo=/collections/An_integrated_single-cell_transcriptomic_dataset_for1_non-small_cell_lung_cancer/6222221.

Regarding our proteomic analysis of activated mouse CD8+ T cells, the MS proteomics data have been deposited to the ProteomeXchange Consortium via the PRIDE (46) partner repository with the dataset identifier PXD050529 (https://www.ebi.ac.uk/pride/archive?keyword=PXD050529&sortDirection=DESC&page=0&pageSize=20). For the mouse NEDD8 IP data, the MS data have been deposited to the ProteomeXchange Consortium via the PRIDE partner repository with the dataset identifier PXD050527 and 10.6019/PXD050527 (https://www.ebi.ac.uk/pride/archive?keyword=PXD050527&sortDirection=DESC&page=0&pageSize=20). Any remaining information can be obtained from the corresponding author upon request.

The R code used for single-cell data reanalysis can be found at https://github.com/eprieto012/NEDD8_APLab.

Differentiation of TILs into effector memory phenotype promotes NEDDylation

To explore the NEDDylation state of immune cells within tumors, we first carried out an in silico approach to analyze publicly available scRNA-seq data from patients with lung cancer. We first examined the expression of NEDD8 in tumor-associated immune and nonimmune populations using the scRNA-seq lung cancer atlas generated by Prazanowska and Lim (45), and we found expression of NEDD8 in all tested populations (Fig. 1A). We then compared NEDD8 expression in CD8+ TIL subpopulations, classified according to their differentiation status using ProjectTILs (47). Higher expression of NEDD8 was observed in effector memory compared with naïve-like CD8+ T cells in three representative datasets from different cancer types: lung cancer (48), melanoma (49), and hepatocarcinoma (Fig. 1B; ref. 50). Based on these findings, we assessed expression of NEDD8 in LLC mouse subcutaneous tumors. Immunofluorescence images in Fig. 1C showed that NEDD8 was expressed in the intracellular compartment of CD3+ TILs. Subsequently, we carried out in vitro experiments to confirm the expression of NEDD8 in murine and human CD8+ T cells, isolated from the spleen and blood of healthy donors, respectively. We found that activation of both mouse (Fig. 1D) and human (Fig. 1E) CD8+ T cells promoted the expression of NEDD8 transcripts. Activated CD8+ T cells also upregulated the expression of UBE2M, a key enzyme of the NEDDylation cycle, whereas NAE1 levels remained constitutive. This transcriptional regulation was accompanied by an enrichment of NEDDylation at the protein level, as evidenced by the accumulation of NEDDylated cullins and free NEDD8 in both mouse and human activated CD8+ T cells (Fig. 1F and G). The protein levels of NAE1 and UBC12 (UBE2M) were also increased upon activation (Fig. 1F and G; Supplementary Fig. S1A and S1B). In addition, stimulation of TCR signaling with an anti-CD3 was sufficient to upregulate NEDDylated proteins and NAE1 (Supplementary Fig. S1C–S1E).

Figure 1.

NEDD8 expression is upregulated during the activation and differentiation into effector memory TILs. A, Uniform Manifold Approximation and Projection (UMAP) plot of immune and nonimmune cell populations generated by analyzing publicly available scRNA-seq data of patients with lung cancer (left). Violin plots showing NEDD8 expression in the different immune and nonimmune cell populations associated with lung cancer microenvironment (right). B, Dot plots showing the mean expression levels of NEDD8 in 5-cell interactions (n = 50) across the indicated types of cancer and phenotypes. Cells with zero expression of NEDD8 were not considered in this analysis. C, Representative immunofluorescence images showing NEDD8 expression in mouse CD3+ TILs (n = 3, scale bar = 10 μm; scale bar for zoomed image = 2 μm). D, Relative RNA expression levels of NEDD8, Ube2m, and Nae1 in mouse primary CD8+ T cells upon activation with anti-CD3 and anti-CD28 antibodies, at the indicated timepoints, measured by RT-qPCR using Rplp0 as housekeeping gene (n = 3, each sample was assessed in triplicate). E, Relative RNA expression levels of NEDD8, UBE2M, and NAE1 in human primary CD8+ T cells upon activation with anti-CD3 and anti-CD28 antibodies, at the specified timepoints, measured by RT-qPCR, using 18S as housekeeping gene (n = 4, two independent experiments, each sample was assessed in triplicate). F, Western blot showing the expression of NEDDylated cullins, NAE1, free NEDD8, β-actin, and UBC12 (UBE2M) in mouse CD8+ T cells, under the described activation conditions (n = 3, two independent experiments). The same membrane was sequentially probed and developed after each incubation with the following antibodies: 1°, anti-NAE1 (top half membrane)/anti-UBC12 (bottom half membrane); 2°, anti–β-actin; 3°, anti-NEDD8. The loading control was run on the same blot as the experimental samples. G, Western blot showing the expression of NEDDylated cullins, NAE1, free NEDD8, β-actin, and UBC12 (UBE2M) in human primary CD8+ T cells under the indicated activation conditions (n = 3, two independent experiments). Two separate membranes were used. The first membrane was sequentially probed and developed after each incubation with the following antibodies: 1°, anti-NAE1; 2°, anti-NEDD8; and 3°, anti–β-actin. A second membrane, loaded with the same protein samples, was used to detect: 1°, UBC12; 2°, β-actin. In both membranes, the loading control was run on the same blot as the experimental samples. Data represented as the mean ± SEM, and analyzed using an unpaired t test. *, P ≤ 0.05; ***, P ≤ 0.001; ****, P ≤ 0.0001.

Figure 1.

NEDD8 expression is upregulated during the activation and differentiation into effector memory TILs. A, Uniform Manifold Approximation and Projection (UMAP) plot of immune and nonimmune cell populations generated by analyzing publicly available scRNA-seq data of patients with lung cancer (left). Violin plots showing NEDD8 expression in the different immune and nonimmune cell populations associated with lung cancer microenvironment (right). B, Dot plots showing the mean expression levels of NEDD8 in 5-cell interactions (n = 50) across the indicated types of cancer and phenotypes. Cells with zero expression of NEDD8 were not considered in this analysis. C, Representative immunofluorescence images showing NEDD8 expression in mouse CD3+ TILs (n = 3, scale bar = 10 μm; scale bar for zoomed image = 2 μm). D, Relative RNA expression levels of NEDD8, Ube2m, and Nae1 in mouse primary CD8+ T cells upon activation with anti-CD3 and anti-CD28 antibodies, at the indicated timepoints, measured by RT-qPCR using Rplp0 as housekeeping gene (n = 3, each sample was assessed in triplicate). E, Relative RNA expression levels of NEDD8, UBE2M, and NAE1 in human primary CD8+ T cells upon activation with anti-CD3 and anti-CD28 antibodies, at the specified timepoints, measured by RT-qPCR, using 18S as housekeeping gene (n = 4, two independent experiments, each sample was assessed in triplicate). F, Western blot showing the expression of NEDDylated cullins, NAE1, free NEDD8, β-actin, and UBC12 (UBE2M) in mouse CD8+ T cells, under the described activation conditions (n = 3, two independent experiments). The same membrane was sequentially probed and developed after each incubation with the following antibodies: 1°, anti-NAE1 (top half membrane)/anti-UBC12 (bottom half membrane); 2°, anti–β-actin; 3°, anti-NEDD8. The loading control was run on the same blot as the experimental samples. G, Western blot showing the expression of NEDDylated cullins, NAE1, free NEDD8, β-actin, and UBC12 (UBE2M) in human primary CD8+ T cells under the indicated activation conditions (n = 3, two independent experiments). Two separate membranes were used. The first membrane was sequentially probed and developed after each incubation with the following antibodies: 1°, anti-NAE1; 2°, anti-NEDD8; and 3°, anti–β-actin. A second membrane, loaded with the same protein samples, was used to detect: 1°, UBC12; 2°, β-actin. In both membranes, the loading control was run on the same blot as the experimental samples. Data represented as the mean ± SEM, and analyzed using an unpaired t test. *, P ≤ 0.05; ***, P ≤ 0.001; ****, P ≤ 0.0001.

Close modal

Genetic and pharmacologic inhibition of NEDDylation hinders CD8+ T-cell survival and proliferation

To assess the role of NEDDylation in CD8+ T-cell function, we generated a transgenic mouse in which NAE1 was conditionally deleted by the Cre recombinase under the control of the distal Lck promoter (NAE1-KO; Fig. 2A). This led to the specific deletion of NAE1 in T cells after their thymic maturation. We confirmed that the thymus and the spleen of NAE1-KO mice contained normal immune populations (Supplementary Fig. S2A–S2C). However, we did observe a modest reduction in the percentage of CD8+ T cells present in lymph nodes (Supplementary Fig. S2D), suggesting a potential role of NEDDylation in mature and antigen-experienced CD8+ T cells.

Figure 2.

NEDDylation promotes the survival and proliferation of CD8+ T cells. A, Schematic representation of the generation of T cell–specific NAE1-KO mouse model. B, Western blot showing the reduction of NEDDylated cullins as a result of the deletion of NAE1 or the pharmacologic inhibition with MLN4924 in mouse CD8+ T cells activated for 3 days (n = 5, four independent experiments). Two separate membranes were used. The first membrane was sequentially probed and developed after each incubation with the following antibodies: 1°, anti-NEDD8; 2°, anti–β-actin. A second membrane, loaded with the same protein samples, was sequentially probed and developed after each incubation with the following antibodies: 1°, anti-NAE1; 2°, anti–β-actin. In both membranes, the loading control was run on the same blot as the experimental samples. C, Western blot showing the reduction of NEDDylated cullins as a result of the pharmacologic inhibition with MLN4924 in CD8+ T cells at the defined doses and activated for 3 days (n = 5, four independent experiments). The same membrane was sequentially probed and developed after each incubation with the following antibodies: 1°, anti-NAE1; 2°, anti-NEDD8; 3°, anti–β-actin. The loading control was run on the same blot as the experimental samples. D, Percentage of alive (7AAD annexin V) and late apoptotic (7AAD+ annexin V+) mouse CD8+ T cells at day 3 (left) and day 6 (right) of activation (n = 3, unpaired parametric t test). E, Percentage of alive human CD8+ T cells (DAPI-negative) at day 3 (left) and day 6 (right) of activation, treated with MLN4924 at the specified doses and analyzed by flow cytometry (n = 3, ordinary one-way ANOVA). F, Percentage of increase in cell count of mouse CD8+ T cells, at the defined timepoints, in control, NAE1-KO or MLN4924-treated CD8+ T cells (n = 3, unpaired parametric t test). G, Representative histograms showing the proliferation of control, NAE1-KO, or MLN4924-treated mouse CD8+ T cells, measured by a CFSE dilution assay at day 3 of activation (n = 3). H, Division index corresponding to the CFSE dilution assay shown in (G)(n = 3, unpaired parametric t test). I, Western blot showing the protein levels of WEE-1 and α-tubulin in control, NAE1-KO, and MLN4924-treated mouse CD8+ T cells, activated for 3 days (n = 4, two independent experiments). The same membrane was sequentially probed and developed after each incubation with the following antibodies: 1°, anti-WEE-1; 2°, anti-β-ACTIN. The loading control was run on the same blot as the experimental samples. J, Representative histograms showing the proliferation of control and MLN4924-treated human CD8+ T cells, measured by a CFSE dilution assay at day 5 of activation (n = 4). K, Division index corresponding to the CFSE dilution assay shown in (I) (n = 4, ordinary one-way ANOVA). L, CD69 expression measured by flow cytometry (gMFI) at day 3 of activation in control (n = 13) and NAE1-KO (n = 14) CD8+ T cells (four independent experiments, unpaired parametric t test. M, CD69 expression measured by flow cytometry (gMFI) 12 hours after activation in control and NAE1-KO CD8+ T cells (ordinary one-way ANOVA. N, NF-κB activity measured in Jurkat TPR cells treated with MLN4924 at the indicated doses, upon activation with anti-CD3 and anti-CD28 for 1 day, and analyzed by flow cytometry (n = 6, two independent experiments, ordinary one-way ANOVA). O, NFAT activity measured in Jurkat TPR cells in the same conditions as in (J) (n = 6, two independent experiments, ordinary one-way ANOVA). Data are represented as the mean ± SEM. *, P ≤ 0.05; **, P ≤ 0.01; ***, P ≤ 0.001; ****, P ≤ 0.0001. gMFI, geometric mean fluorescence intensity.

Figure 2.

NEDDylation promotes the survival and proliferation of CD8+ T cells. A, Schematic representation of the generation of T cell–specific NAE1-KO mouse model. B, Western blot showing the reduction of NEDDylated cullins as a result of the deletion of NAE1 or the pharmacologic inhibition with MLN4924 in mouse CD8+ T cells activated for 3 days (n = 5, four independent experiments). Two separate membranes were used. The first membrane was sequentially probed and developed after each incubation with the following antibodies: 1°, anti-NEDD8; 2°, anti–β-actin. A second membrane, loaded with the same protein samples, was sequentially probed and developed after each incubation with the following antibodies: 1°, anti-NAE1; 2°, anti–β-actin. In both membranes, the loading control was run on the same blot as the experimental samples. C, Western blot showing the reduction of NEDDylated cullins as a result of the pharmacologic inhibition with MLN4924 in CD8+ T cells at the defined doses and activated for 3 days (n = 5, four independent experiments). The same membrane was sequentially probed and developed after each incubation with the following antibodies: 1°, anti-NAE1; 2°, anti-NEDD8; 3°, anti–β-actin. The loading control was run on the same blot as the experimental samples. D, Percentage of alive (7AAD annexin V) and late apoptotic (7AAD+ annexin V+) mouse CD8+ T cells at day 3 (left) and day 6 (right) of activation (n = 3, unpaired parametric t test). E, Percentage of alive human CD8+ T cells (DAPI-negative) at day 3 (left) and day 6 (right) of activation, treated with MLN4924 at the specified doses and analyzed by flow cytometry (n = 3, ordinary one-way ANOVA). F, Percentage of increase in cell count of mouse CD8+ T cells, at the defined timepoints, in control, NAE1-KO or MLN4924-treated CD8+ T cells (n = 3, unpaired parametric t test). G, Representative histograms showing the proliferation of control, NAE1-KO, or MLN4924-treated mouse CD8+ T cells, measured by a CFSE dilution assay at day 3 of activation (n = 3). H, Division index corresponding to the CFSE dilution assay shown in (G)(n = 3, unpaired parametric t test). I, Western blot showing the protein levels of WEE-1 and α-tubulin in control, NAE1-KO, and MLN4924-treated mouse CD8+ T cells, activated for 3 days (n = 4, two independent experiments). The same membrane was sequentially probed and developed after each incubation with the following antibodies: 1°, anti-WEE-1; 2°, anti-β-ACTIN. The loading control was run on the same blot as the experimental samples. J, Representative histograms showing the proliferation of control and MLN4924-treated human CD8+ T cells, measured by a CFSE dilution assay at day 5 of activation (n = 4). K, Division index corresponding to the CFSE dilution assay shown in (I) (n = 4, ordinary one-way ANOVA). L, CD69 expression measured by flow cytometry (gMFI) at day 3 of activation in control (n = 13) and NAE1-KO (n = 14) CD8+ T cells (four independent experiments, unpaired parametric t test. M, CD69 expression measured by flow cytometry (gMFI) 12 hours after activation in control and NAE1-KO CD8+ T cells (ordinary one-way ANOVA. N, NF-κB activity measured in Jurkat TPR cells treated with MLN4924 at the indicated doses, upon activation with anti-CD3 and anti-CD28 for 1 day, and analyzed by flow cytometry (n = 6, two independent experiments, ordinary one-way ANOVA). O, NFAT activity measured in Jurkat TPR cells in the same conditions as in (J) (n = 6, two independent experiments, ordinary one-way ANOVA). Data are represented as the mean ± SEM. *, P ≤ 0.05; **, P ≤ 0.01; ***, P ≤ 0.001; ****, P ≤ 0.0001. gMFI, geometric mean fluorescence intensity.

Close modal

We confirmed that the deletion of NAE1 in CD8+ T cells resulted in a reduction of NEDDylated cullins upon activation (Fig. 2B). Similarly, treatment with the pharmacologic inhibitor of NAE, MLN4924, also reduced the amount of NEDDylated cullins in both mouse (Fig. 2B) and human (Fig. 2C) CD8+ T cells. Thereafter, we measured the viability and proliferation of mouse CD8+ T cells with defective NEDDylation, either isolated from NAE1-KO mice or treated with MLN4924. Both conditions resulted in a decrease in viability and an increase in late apoptotic cells, as measured by 7AAD and annexin V staining (Fig. 2D; Supplementary Fig. S3A). Similar reductions in viability were observed in human CD8+ T cells treated with MLN4924 (Fig. 2E; Supplementary Fig. S3B). Mouse CD8+ T cells with deficient NEDDylation had a defect in proliferation, measured by cell counting (Fig. 2F) and by CFSE dilution (Fig. 2G and H), potentially resulting from the accumulation of cell-cycle regulators such us WEE-1, P21, and P27 (Fig. 2I; Supplementary Fig. S3C–S3E), as previously described for other cell types (51). A similar proliferation impairment was observed when human CD8+ T cells were treated with MLN4924 (Fig. 2J and K).

The reduced viability and proliferation of NEDDylation-deficient CD8+ T cells may result from attenuated T-cell activation, as indicated by reduced CD69 expression in NAE1-KO CD8+ T cells (Fig. 2L) and in human CD8+ T cells treated with increasing doses of MLN4924 (Fig. 2M). We also measured the activity of the NF-κB and NFAT signaling pathways in Jurkat TPR cells (5254). NF-κB and NFAT are important transcriptional regulators for CD8+ T-cell survival and proliferation (5557), and their activity was reduced when Jurkat TPR cells were treated with MLN4924 (Fig. 2N and O).

NEDDylation supports T cell–mediated antitumor immunity

We next sought to validate whether these defects had functional consequences in vivo. Because the role of NEDDylation in T cell–mediated antitumor immunity is poorly described, we decided to investigate the impact of NEDDylation in the ability of T cells to fight against solid tumors. To this end, we subcutaneously implanted LLC cells in control or NAE-1-KO mice. We observed that dysfunctional NEDDylation in T cells reduced their capacity to control tumor growth (Fig. 3A and B). In this context, the amount of CD4+ (Fig. 3C) and CD8+ T cells (Fig. 3D) found per milligram of tumor was severely reduced in NAE1-KO mice. A decrease in the amount of NK cells (NK-1.1+) per milligram of tumor was also observed, whereas the amounts of intratumoral B lymphocytes (CD19+) and myeloid cells (CD11b+) were not affected (Supplementary Fig. S4A). In addition, NAE1-KO CD8+ TILs showed a less differentiated phenotype, as measured by the expression of CD44 and CD62L (Supplementary Fig. S4B) and lower levels of LAG-3, PD-1, CD137, and TIM-3 (Supplementary Fig. S4C and S4D). Furthermore, the percentage of intratumoral Ki-67+ T cells was reduced in NAE1-KO mice (Supplementary Fig. S4E and S4F). Immunophenotyping of secondary lymphoid organs revealed no differences in immune-cell composition in the spleen (Supplementary Fig. S4G) and a reduction in the proportion of CD8+ T cells in the tumor-draining lymph nodes (Supplementary Fig. S4H).

Figure 3.

Deletion of NAE1 accelerates tumor growth and decreases the number of TILs, impairing their differentiation into effector CD8+ cells. A, Mean tumor volume of control (n = 14) and NAE1-KO (n = 11) mice after the subcutaneous injection of the LLC tumor cell line (three independent experiments, unpaired nonparametric Mann–Whitney test). B, Tumor volume measurement of control (n = 14) and NAE1-KO (n = 11) mice at endpoint day after subcutaneous injection of the LLC tumor cell line (unpaired nonparametric Mann–Whitney test). C, Absolute number of CD4+ T cells per milligram of tumor at endpoint day in control (n = 14) and NAE1-KO (n = 11) mice after subcutaneous injection of the LLC tumor cell line, analyzed by flow cytometry (unpaired nonparametric Mann–Whitney test). D, Absolute number of CD8+ T cells per milligram of tumor, in the same conditions as in (C). E, Mean tumor volume of control (n = 11) and NAE1-KO (n = 12) mice after the subcutaneous injection of YUMMER 1.7-GFP tumor cell line (two independent experiments, unpaired nonparametric Mann–Whitney test). F, Percentage of T-bet+ CD4+ T cells (unpaired nonparametric Mann–Whitney test). G, Intratumoral CD8+ T-cell phenotype of control (n = 4) and NAE1-KO (n = 6) mice analyzed by flow cytometry (effector: CD44+ CD62L; central memory: CD44+ CD62L+; naïve: CD44 CD62L+, unpaired nonparametric Mann–Whitney test). H, Percentage of intratumoral CD8+ T cells triple-positive for PD-1, TIM-3, and TIGIT (unpaired nonparametric Mann–Whitney test). I, Percentage of intratumoral CD8+ T cells double-positive for CD39 and TIM-3 (unpaired nonparametric Mann–Whitney test). J, Percentage of intratumoral CD8+ T cells double-positive for CD44high and PD-1 (unpaired nonparametric Mann–Whitney test). K, Percentage of TCF7+ cells in intratumoral CD8+ T cells double-positive for CD44high and PD-1 (unpaired nonparametric Mann–Whitney test). L, Percentage of granzyme B+ intratumoral CD8+ T cells (unpaired nonparametric Mann–Whitney test). M, Individual tumor development of untreated (discontinuous black, n = 6), control-OT-I (gray, n = 9), or NAE1-KO-OT-I (red, n = 8) treated mice. Tumor growth was measured at the indicated timepoints (two independent experiments). N, Mean tumor development of untreated (n = 6), control-OT-I (n = 9), or NAE1-KO-OT-I (n = 8)–treated mice between OT-I injection day and day 20, when the first mouse reached endpoint (two independent experiments, one-way ANOVA–Kruskal–Wallis test). O, Mean tumor volume measured in the previously indicated groups at day 20 (one-way ANOVA–Kruskal–Wallis test). P, Kaplan–Meier survival analysis of the specified groups (log-rank Mantel–Cox test). Q, Western Blot showing the expression of NEDDylated proteins in control (hAAVS1 KO as safe harbor gene) and SENP8 KO CD8+ T cells (n = 4, two independent experiments). The same membrane was sequentially probed and developed after each incubation with the following antibodies: 1°, anti-NEDD8; 2°, anti–β-actin; 3°, anti-SENP8. The loading control was run on the same blot as the experimental samples. R, Percentage of lysis of RAMOS-ZsGreen cells cocultured with human SENP8-KO CD8+ CAR T cells. Cytotoxicity assay was performed for 24 hours at the indicated effector (T cell) to target (tumor cell) ratios (E:T ratios). (n = 2, two-way ANOVA). Data are represented as the mean ± SEM. *, P ≤ 0.05; **, P ≤ 0.01; ***, P ≤ 0.001; ****, P ≤ 0.0001; ns, not significant.

Figure 3.

Deletion of NAE1 accelerates tumor growth and decreases the number of TILs, impairing their differentiation into effector CD8+ cells. A, Mean tumor volume of control (n = 14) and NAE1-KO (n = 11) mice after the subcutaneous injection of the LLC tumor cell line (three independent experiments, unpaired nonparametric Mann–Whitney test). B, Tumor volume measurement of control (n = 14) and NAE1-KO (n = 11) mice at endpoint day after subcutaneous injection of the LLC tumor cell line (unpaired nonparametric Mann–Whitney test). C, Absolute number of CD4+ T cells per milligram of tumor at endpoint day in control (n = 14) and NAE1-KO (n = 11) mice after subcutaneous injection of the LLC tumor cell line, analyzed by flow cytometry (unpaired nonparametric Mann–Whitney test). D, Absolute number of CD8+ T cells per milligram of tumor, in the same conditions as in (C). E, Mean tumor volume of control (n = 11) and NAE1-KO (n = 12) mice after the subcutaneous injection of YUMMER 1.7-GFP tumor cell line (two independent experiments, unpaired nonparametric Mann–Whitney test). F, Percentage of T-bet+ CD4+ T cells (unpaired nonparametric Mann–Whitney test). G, Intratumoral CD8+ T-cell phenotype of control (n = 4) and NAE1-KO (n = 6) mice analyzed by flow cytometry (effector: CD44+ CD62L; central memory: CD44+ CD62L+; naïve: CD44 CD62L+, unpaired nonparametric Mann–Whitney test). H, Percentage of intratumoral CD8+ T cells triple-positive for PD-1, TIM-3, and TIGIT (unpaired nonparametric Mann–Whitney test). I, Percentage of intratumoral CD8+ T cells double-positive for CD39 and TIM-3 (unpaired nonparametric Mann–Whitney test). J, Percentage of intratumoral CD8+ T cells double-positive for CD44high and PD-1 (unpaired nonparametric Mann–Whitney test). K, Percentage of TCF7+ cells in intratumoral CD8+ T cells double-positive for CD44high and PD-1 (unpaired nonparametric Mann–Whitney test). L, Percentage of granzyme B+ intratumoral CD8+ T cells (unpaired nonparametric Mann–Whitney test). M, Individual tumor development of untreated (discontinuous black, n = 6), control-OT-I (gray, n = 9), or NAE1-KO-OT-I (red, n = 8) treated mice. Tumor growth was measured at the indicated timepoints (two independent experiments). N, Mean tumor development of untreated (n = 6), control-OT-I (n = 9), or NAE1-KO-OT-I (n = 8)–treated mice between OT-I injection day and day 20, when the first mouse reached endpoint (two independent experiments, one-way ANOVA–Kruskal–Wallis test). O, Mean tumor volume measured in the previously indicated groups at day 20 (one-way ANOVA–Kruskal–Wallis test). P, Kaplan–Meier survival analysis of the specified groups (log-rank Mantel–Cox test). Q, Western Blot showing the expression of NEDDylated proteins in control (hAAVS1 KO as safe harbor gene) and SENP8 KO CD8+ T cells (n = 4, two independent experiments). The same membrane was sequentially probed and developed after each incubation with the following antibodies: 1°, anti-NEDD8; 2°, anti–β-actin; 3°, anti-SENP8. The loading control was run on the same blot as the experimental samples. R, Percentage of lysis of RAMOS-ZsGreen cells cocultured with human SENP8-KO CD8+ CAR T cells. Cytotoxicity assay was performed for 24 hours at the indicated effector (T cell) to target (tumor cell) ratios (E:T ratios). (n = 2, two-way ANOVA). Data are represented as the mean ± SEM. *, P ≤ 0.05; **, P ≤ 0.01; ***, P ≤ 0.001; ****, P ≤ 0.0001; ns, not significant.

Close modal

We performed additional in vivo experiments with the melanoma cell line YUMMER 1.7-GFP, which is more immunogenic than the LLC cell line (58, 59). Again, tumor growth was accelerated in NAE1-KO mice (Fig. 3E; Supplementary Fig. S4I), and the amount of CD4+ and CD8+ T cells was drastically reduced per milligram of tumor (Supplementary Fig. S4J and S4K). Analyses of the composition of the NAE1-KO TILs revealed a reduction in the percentage of T-bet+CD4+ T cells (Fig. 3F) and CD44+CD62L-CD8+ T cells (Fig. 3G), suggesting an impairment of the CD4+ Th1 phenotype and CD8+ T-cell effector differentiation, respectively. Within the CD8+ T-cell compartment, NAE1 deletion resulted in a decrease in the percentage of PD-1+TIM-3+TIGIT+ triple-positive T cells (Fig. 3H), CD39+TIM-3+ double-positive T cells (Fig. 3I), and CD44highPD-1+ T cells (Fig. 3J). Whereas control cells expressed higher quantities of these exhaustion markers, TCF7 expression was increased in CD44highPD-1+ NAE1-KO CD8+ T cells, indicating a higher generation of progenitor exhausted cells (Fig. 3K). Overall, the ability of NAE1-KO CD8+ T cells to produce granzyme B was significantly impaired (Fig. 3L), contributing to their diminished antitumor functionality, which parallels their reduced differentiation capability. Next, we evaluated the role of NEDDylation in T-cell exhaustion using the protocol described by Scharping and colleagues (43), which involves repeated in vitro stimulations to induce exhaustion in CD8+ T cells. The levels of NEDDylated proteins were significantly elevated in exhausted CD8+ T cells (Supplementary Fig. S5A and S5B). Furthermore, whereas the percentage of exhausted cells was comparable between control and NAE1-KO CD8+ T cells (Supplementary Fig. S5C), the functional capability of NAE1-KO cells was diminished, as indicated by a reduction in the production of IFNγ (Supplementary Fig. S5D).

To further explore this phenotype, we took advantage of the OT-I antigen–specific model, in which the role of CD8+ T cells is critical. For this purpose, we crossed NAE1-KO mice with OT-I mice to generate control-OT-I mice and NAE1-KO-OT-I mice. We then transferred activated control-OT-I or NAE1-KO-OT-I CD8+ T cells into LLC-OVA tumor-bearing mice and analyzed tumor progression and mice survival. OT-I cells were sorted to isolate and inject only alive cells. Whereas transfer of control-OT-I CD8+ T cells resulted in a delay in tumor growth, transfer of NAE1-KO-OT-I cells showed a diminished therapeutic activity and less ability to control tumor progression (Fig. 3M–O). Consequently, the survival rate of mice infused with NAE1-KO-OT-I cells was significantly reduced in comparison with mice infused with control-OT-I cells (Fig. 3P). We quantified the levels of OT-I CD8+ T-cell infiltration taking advantage of the congenic marker CD45.1+, finding that CD45.1+ NAE1-KO-OT-I cells infiltrated the tumors of CD45.2+ recipient mice in similar quantities to the CD45.1+ control-OT-I cells (Supplementary Fig. S6A). At this point, tumor weight was significantly greater in mice receiving NAE1-KO-OT-I cells (Supplementary Fig. S6B). We also found that the in vitro cytotoxic function of NAE1-KO-OT-I cells was reduced by coculturing control-OT-I and NAE1-KO-OT-I cells with B16-luciferase-OVA target cells (Supplementary Fig. S6C and S6D).

Given the consistent observation that NAE1-KO CD8+ T cells displayed a reduced antitumor function, we next explored whether upregulating NEDDylation could have the opposite effect. For this, we developed an approach to enhance NEDDylation in CD8+ T cells based on knocking out the deNEDDylase SENP8, which specifically removes NEDD8 from non-cullin substrates (60, 61). As a result, the amount of NEDDylated proteins was increased in human CD8+ T cells (Fig. 3Q). To explore its antitumor effects, we generated SENP8-KO anti-CD19 CAR CD8+ T cells and cocultured them with CD19-expressing RAMOS target cells. Knocking out SENP8 in CAR T cells enhanced their cytotoxic capability (Fig. 3R). In summary, these findings show the importance of NEDDylation in the antitumor function of CD8+ T cells.

Key metabolic pathways are impaired in NAE1-KO or MLN4924-treated CD8+ T cells

Our in vivo assays revealed that NEDDylation has an important effect on CD8+ T-cell function. To obtain a detailed perspective of the effects of NEDD8 in CD8+ T cells, we performed a proteomic analysis in NAE1-KO CD8+ T cells and MLN4924-treated CD8+ T cells, compared with control cells. We identified many significantly dysregulated proteins (Supplementary Table S2). These were analyzed by IPA, revealing several dysregulated pathways upon inhibition of NEDDylation (Fig. 4A; Supplementary Tables S2 and S3), including glycolysis, the HIF-1α pathway, and oxidative phosphorylation (Fig. 4B). Additionally, other identified pathways were the TCR, migration, and MYC signaling pathways (Supplementary Fig. S7A–S7C). To functionally validate these findings, we evaluated the metabolic activity of NAE1-KO CD8+ T cells by measuring the extracellular acidification rate (Fig. 4C) and oxygen consumption rate (Fig. 4D). NAE1-KO CD8+ T cells had lower basal glycolysis (Fig. 4E), compensatory glycolysis (Fig. 4F), and basal respiration (Fig. 4G). Spare respiratory capacity and maximal respiration were also impaired (Supplementary Fig. S7D and S7E). A similar phenotype was observed in MLN4924-treated human CD8+ T cells (Fig. 4H and I), leading to lower basal glycolysis (Fig. 4J), compensatory glycolysis (Fig. 4K), and basal respiration (Fig. 4L). To gain further insight into the link between NEDDylation, metabolism, and cytotoxic capability, we assessed granzyme B production in control and NAE1-KO CD8+ T cells that had been treated with galactose to restrict their glycolysis. A second strategy to inhibit glycolysis was treatment with 2-DG, a stronger inhibitor of the glycolytic pathway (bioRxiv.2020.02.05.935627). Under standard culture conditions, NAE1-KO CD8+ T cells exhibited a diminished ability to produce granzyme B (Supplementary Fig. S7F). Glycolysis inhibition with galactose decreased granzyme B production in control cells but had no effect on NAE1-KO CD8+ T cells. Consistently, 2-DG treatment compromised granzyme B production in control cells to the levels observed in NAE1-KO cells (Supplementary Fig. S7F), suggesting that the observed reduction of granzyme B production in NAE1-KO T cells could be induced by their metabolic impairment.

Figure 4.

Deletion or inhibition of NAE1 in activated CD8+ T cells impairs key metabolic pathways. A, Impaired pathways in NAE1-KO and MLN4924-treated CD8+ T cells activated for 3 days, in comparison with control cells. Z-score and P values are obtained by the IPA software (n = 4 per condition, Fisher exact test). B, Heatmap showing row-normalized dysregulated proteins in metabolic pathways (glycolysis, HIF-1α signaling, and oxidative phosphorylation) comparing NAE1-KO and MLN4924-treated CD8+ T cells with control CD8+ T cells. (n = 4 per condition). C, Representative scheme for extracellular acidification rate (ECAR) measurement in control and NAE1-KO activated mouse CD8+ T cells, activated for 3 days and measured using Agilent Seahorse XF Cell Glycolytic Rate Assay Kit (one mouse represented per condition, total n = 3). D, Representative scheme for oxygen consumption rate (OCR) measurement in control and NAE1-KO mouse CD8+ T cells, activated for 3 days and measured with Agilent Seahorse XF Cell Mito Stress Test Kit (one mouse represented per condition, total n = 3). E, Basal glycolysis and (F) compensatory glycolysis expressed as the glycolytic proton efflux rate (GlycoPER, unpaired parametric t test). G, Basal respiration, expressed as OCR measurement (unpaired parametric t tests). H, Representative scheme for ECAR measurement in control and MLN4924-treated human CD8+ T cells, activated for 3 days and measured using Agilent Seahorse XF Cell Glycolytic Rate Assay Kit (one healthy donor represented per condition, total n = 4 from two independent experiments). I, Representative scheme for OCR measurement in control and MLN4924-treated activated human CD8+ T cells, activated for 3 days and measured using Agilent Seahorse XF Cell Mito Stress Test Kit (one healthy donor represented per condition, total n = 3). J, Basal glycolysis and (K) compensatory glycolysis expressed as the GlycoPER (paired parametric t test). L, Basal respiration expressed OCR measurement (paired parametric t test). Data are represented as the mean ± SEM. *, P ≤ 0.05; **, P ≤ 0.01; ***, P ≤ 0.001. Rot/AA, rotenone and antimycin A.

Figure 4.

Deletion or inhibition of NAE1 in activated CD8+ T cells impairs key metabolic pathways. A, Impaired pathways in NAE1-KO and MLN4924-treated CD8+ T cells activated for 3 days, in comparison with control cells. Z-score and P values are obtained by the IPA software (n = 4 per condition, Fisher exact test). B, Heatmap showing row-normalized dysregulated proteins in metabolic pathways (glycolysis, HIF-1α signaling, and oxidative phosphorylation) comparing NAE1-KO and MLN4924-treated CD8+ T cells with control CD8+ T cells. (n = 4 per condition). C, Representative scheme for extracellular acidification rate (ECAR) measurement in control and NAE1-KO activated mouse CD8+ T cells, activated for 3 days and measured using Agilent Seahorse XF Cell Glycolytic Rate Assay Kit (one mouse represented per condition, total n = 3). D, Representative scheme for oxygen consumption rate (OCR) measurement in control and NAE1-KO mouse CD8+ T cells, activated for 3 days and measured with Agilent Seahorse XF Cell Mito Stress Test Kit (one mouse represented per condition, total n = 3). E, Basal glycolysis and (F) compensatory glycolysis expressed as the glycolytic proton efflux rate (GlycoPER, unpaired parametric t test). G, Basal respiration, expressed as OCR measurement (unpaired parametric t tests). H, Representative scheme for ECAR measurement in control and MLN4924-treated human CD8+ T cells, activated for 3 days and measured using Agilent Seahorse XF Cell Glycolytic Rate Assay Kit (one healthy donor represented per condition, total n = 4 from two independent experiments). I, Representative scheme for OCR measurement in control and MLN4924-treated activated human CD8+ T cells, activated for 3 days and measured using Agilent Seahorse XF Cell Mito Stress Test Kit (one healthy donor represented per condition, total n = 3). J, Basal glycolysis and (K) compensatory glycolysis expressed as the GlycoPER (paired parametric t test). L, Basal respiration expressed OCR measurement (paired parametric t test). Data are represented as the mean ± SEM. *, P ≤ 0.05; **, P ≤ 0.01; ***, P ≤ 0.001. Rot/AA, rotenone and antimycin A.

Close modal

The HIF response can directly influence metabolic T-cell adaptations in certain conditions (62). We confirmed that Hif1α mRNA levels were similar in both NAE1-KO and MLN4924-treated CD8+ T cells compared with control cells (Supplementary Fig. S7G). At the protein level, HIF-1α accumulated in both models of NEDDylation deficiency (Supplementary Fig. S7H), an effect previously documented in other cell lines (63, 64). However, this accumulation did not correlate with higher HIF-1α transcriptional activity. In fact, NAE1-KO CD8+ T cells exhibited reduced expression of HIF-1α target genes (Ldha, Glut1, Pdk1, and Pgk1; Supplementary Fig. S7I; ref. 65), confirming the impaired HIF-1α activity in NAE1-KO CD8+ T cells shown in Fig. 4A.

Glycolytic enzymes are targets of NEDD8 in activated CD8+ T cells

We next focused on the identification of NEDDylated proteins that could mediate the observed CD8+ T-cell phenotype. To this end, we performed an IP of NEDD8 in activated mouse CD8+ T cells, followed by MS analysis (Fig. 5A). Numerous proteins were identified as NEDD8 targets (Supplementary Table S4) and analyzed using the Gene Ontology database (66). A variety of cellular metabolic processes were found to be enriched (Fig. 5B). Glycolysis was one of the top scoring pathways, and key glycolytic enzymes were found in the list of NEDD8 targets, such as LDHA, α-enolase (ENO1), and hexokinase 1 (HK-1; ref. 67). Given the central role of glycolysis in controlling T-cell function, we carried out an IP assay in activated human CD8+ T cells, in which we validated LDHA, ENO1, and HK-1 as targets of NEDD8 (Fig. 5C). Furthermore, LDHA protein expression was downregulated in a dose-dependent manner (Fig. 5D) in MLN4924-treated human CD8+ T cells, whereas the levels of ENO1 and HK-1 remained unchanged. This reduction in LDHA expression was also observed in NAE1-KO CD8+ T cells and in MLN4924-treated mouse CD8+ T cells (Fig. 5E). Densitometric quantifications are shown in Supplementary Fig. S8A and S8B. We observed that inhibition of NEDDylation led to a transcriptional impairment of LDHA, as evidenced by reduced transcript levels in both human (Fig. 5F) and mouse (Fig. 5G) NEDDylation-deficient CD8+ T cells. This transcriptional downregulation was accompanied by a decrease in LDHA activity (Fig. 5H), suggesting a link between defective NEDDylation and glycolytic dysfunction.

Figure 5.

NEDD8 binds to three glycolytic enzymes, affecting the transcript levels, the protein expression, and the activity of LDHA. A, Workflow of the process of validation of new NEDD8 targets. B, Dot plot representing enriched biological processes in the list of NEDD8 targets obtained from proteomics analysis (n = 2, two independent experiments). Data obtained from the Gene Ontology (GO) database. C, Western blot showing immunoprecipitated proteins by the NEDD8 antibody in human CD8+ T cells activated for 3 days. Input and IgG controls are shown (n = 3, three independent experiments). NEDDylated cullins are shown as a positive control. Two separate membranes were used. The first membrane was sequentially probed and developed after each incubation with the following antibodies: 1°, anti-NEDD8 (top half)/anti-ENO1 (bottom half); 2°, anti-HK-1 (top half). A second membrane, loaded with the same protein samples, was probed and developed after incubation with anti-LDH antibody. D, Western blot showing the expression of LDHA, ENO1, and HK-1 in activated human CD8+ T cells treated with MLN4924 at the indicated doses for 3 days (n = 3, two independent experiments). The same membrane was sequentially probed and developed after each incubation with the following antibodies: 1°, anti-HK-1 (top half)/anti-LDHA (bottom half); 2°, anti-NEDD8 (top half)/anti-ENO1 (bottom half); 3°, anti–β-ACTIN. The loading control was run on the same blot as the experimental samples. E, Western blot showing the expression of LDHA in control, NAE1-KO, and MLN4924-treated activated mouse CD8+ T cells after 3 days of culture (n = 5, two independent experiments). The same membrane was sequentially probed and developed after each incubation with the following antibodies: 1°, anti-NAE1; 2°, anti-NEDD8; 3°, anti–β-actin; 4°, anti-LDHA. The loading control was run on the same blot as the experimental samples. F, Relative mRNA expression levels of LDHA in activated human primary CD8+ T cells measured by RT-qPCR using 18S as housekeeping gene after 2 days of culture (n = 4, each sample was assessed in triplicate and analyzed using ordinary one-way ANOVA). G, Relative mRNA expression levels of LDHA in activated mouse primary CD8+ T cells measured by RT-qPCR using 36b4 as housekeeping gene after 2 days of culture (n = 4, each sample was assessed in triplicate and analyzed using an unpaired parametric t test). H, LDH activity measured in supernatants of control, MLN4924-treated, and NAE1-KO CD8+ T cells after 3 days of culture. (n = 7, two independent experiments; unpaired parametric t test). Data are represented as the mean ± SEM. *, P ≤ 0.05; **, P ≤ 0.01; ***, P ≤ 0.001.

Figure 5.

NEDD8 binds to three glycolytic enzymes, affecting the transcript levels, the protein expression, and the activity of LDHA. A, Workflow of the process of validation of new NEDD8 targets. B, Dot plot representing enriched biological processes in the list of NEDD8 targets obtained from proteomics analysis (n = 2, two independent experiments). Data obtained from the Gene Ontology (GO) database. C, Western blot showing immunoprecipitated proteins by the NEDD8 antibody in human CD8+ T cells activated for 3 days. Input and IgG controls are shown (n = 3, three independent experiments). NEDDylated cullins are shown as a positive control. Two separate membranes were used. The first membrane was sequentially probed and developed after each incubation with the following antibodies: 1°, anti-NEDD8 (top half)/anti-ENO1 (bottom half); 2°, anti-HK-1 (top half). A second membrane, loaded with the same protein samples, was probed and developed after incubation with anti-LDH antibody. D, Western blot showing the expression of LDHA, ENO1, and HK-1 in activated human CD8+ T cells treated with MLN4924 at the indicated doses for 3 days (n = 3, two independent experiments). The same membrane was sequentially probed and developed after each incubation with the following antibodies: 1°, anti-HK-1 (top half)/anti-LDHA (bottom half); 2°, anti-NEDD8 (top half)/anti-ENO1 (bottom half); 3°, anti–β-ACTIN. The loading control was run on the same blot as the experimental samples. E, Western blot showing the expression of LDHA in control, NAE1-KO, and MLN4924-treated activated mouse CD8+ T cells after 3 days of culture (n = 5, two independent experiments). The same membrane was sequentially probed and developed after each incubation with the following antibodies: 1°, anti-NAE1; 2°, anti-NEDD8; 3°, anti–β-actin; 4°, anti-LDHA. The loading control was run on the same blot as the experimental samples. F, Relative mRNA expression levels of LDHA in activated human primary CD8+ T cells measured by RT-qPCR using 18S as housekeeping gene after 2 days of culture (n = 4, each sample was assessed in triplicate and analyzed using ordinary one-way ANOVA). G, Relative mRNA expression levels of LDHA in activated mouse primary CD8+ T cells measured by RT-qPCR using 36b4 as housekeeping gene after 2 days of culture (n = 4, each sample was assessed in triplicate and analyzed using an unpaired parametric t test). H, LDH activity measured in supernatants of control, MLN4924-treated, and NAE1-KO CD8+ T cells after 3 days of culture. (n = 7, two independent experiments; unpaired parametric t test). Data are represented as the mean ± SEM. *, P ≤ 0.05; **, P ≤ 0.01; ***, P ≤ 0.001.

Close modal

Our study elucidates the pivotal role of NEDDylation, mediated by NAE1, in the regulation of CD8+ T-cell metabolism and antitumor immunity, highlighting its significance beyond the well-documented impact on tumor cell proliferation and survival (68). We have identified that TILs upregulate the NEDDylation pathway as they differentiate into effector memory cells, favoring an adaptive metabolic shift that is essential for their antitumor function. Specifically, we show that both genetic and pharmacologic inhibition of NEDDylation severely impair CD8+ T-cell viability, proliferation, and metabolic pathways, resulting in a reduction of their effector function.

Our findings have implications for the current therapeutic landscape, in which NEDDylation inhibitors, such as MLN4924, are under clinical investigation for cancer treatment. In vivo, MLN4924 is effective in controlling tumor growth and promoting antitumor immunity (33, 34). However, the direct effects of inhibiting NEDDylation in T cells require further understanding. Recent clinical trial shortcomings of MLN4924 against several tumor types have tempered expectations, highlighting the need for further research of this pathway, including its immune modulation properties. For this purpose, the NAE1-KO model provides a tool to elucidate the roles of NEDDylation specifically in CD8+ T cells.

Beyond NAE inhibitors, emerging antitumor approaches include targeting NEDDylation E2 conjugating enzymes such as UBE2M (69). However, inhibiting UBE2M can also compromise CD4+ T-cell function (70) and could similarly impact CD8+ T-cell function given that UBE2M expression is upregulated by activated CD8+ T cells.

In this study, we describe the detrimental effects of NEDDylation inhibition on CD8+ T-cell function, demonstrating a defect in their ability to combat tumor cells. These findings can help to explain the limited efficacy of NEDDylation inhibitors in the clinic. This paradox underscores the complexity of targeting cellular processes that are pivotal for both tumor progression and immune defense mechanisms.

In this context, we show that NEDDylation favors the acquisition of a glycolytic phenotype by T cells. Specifically, we demonstrate that NEDDylation-deficient CD8+ T cells are characterized by reduced LDHA transcript levels, protein expression, and enzymatic activity. This finding aligns with the dysregulation of MYC and HIF1-α signaling pathways, known regulators of LDHA and the associated metabolic reprogramming that drives T-cell activation (65, 71). In fact, metabolic reprogramming is a central feature of the robustness of immune responses (72). In the tumor microenvironment, effector T cells rely on LDHA for their differentiation and function (73), a finding consistent with our observation of impaired differentiation in vivo during tumor-challenge assays.

This mechanism can offer opportunities for therapeutic intervention by modulating the NEDDylation of these glycolytic enzymes and metabolic pathways to enhance the metabolic fitness and antitumor capabilities of T cells. We identified the glycolytic enzymes LDHA, HK-1, and ENO1 as NEDD8 targets in activated CD8+ T cells, suggesting another layer of regulation of metabolism by NEDDylation, similar to that observed with other cell types (74, 75). PTMs are known to regulate multiple members of the same signaling pathway (61), suggesting that NEDDylation could control glycolysis acting on several enzymes of this pathway. Determining whether these enzymes are regulated posttranslationally by NEDD8 is a critical area of investigation that requires further elucidation.

The specific regulation of these NEDD8 targets is of particular interest, and the substrate specificity of E3 ligases can aid in this purpose (76). Identifying the E3 ligases involved in the specific NEDDylation of these substrates can avoid the negative outcomes of overall NEDDylation inhibition and unlock new ways to manipulate protein–protein interactions, such as using proteolysis-targeting chimeras. In a similar manner, the extension of PTM strategies into the field of adoptive cell therapies is a novel approach. For example, engineering of CAR T cells to circumvent CAR ubiquitination enhances their antitumor effectiveness and endurance within the host (77, 78). In this study, we propose an approach wherein knocking out the deNEDDylase SENP8 confers superior antitumor characteristics to CD8+ CAR T cells. This strategy involves upregulating NEDDylation in CD8+ T cells, serving as a mechanism to optimize a variety of adoptive cell therapies, including engineered CAR and TCR approaches. An alternative approach was recently proposed by Liao and colleagues (79), based on the inhibition of cullin 5 in CAR T cells. Notably, adoptive cell therapies provide an opportunity to specifically treat engineered T cells without requiring systemic administration of the tested agent, which can be useful to regulate NEDDylation to keep its beneficial outcomes while avoiding its deleterious effects.

In conclusion, our study highlights NEDDylation as a critical regulatory mechanism in CD8+ T-cell metabolism and antitumor immunity, delineating the dual role of NEDDylation in tumor cells and immune cells, and exploring molecular pathways supporting potential therapeutic strategies that harness this pathway for cancer immunotherapy.

B. Jiménez-Lasheras reports a patent for EP21382780 pending. A. Palazon reports a patent for EP21382780 pending. No disclosures were reported by the other authors.

B. Jiménez-Lasheras: Conceptualization, resources, formal analysis, validation, investigation, visualization, methodology, writing–original draft, writing–review and editing. P. Velasco-Beltrán: Conceptualization, resources, formal analysis, validation, investigation, visualization, methodology, writing–original draft, writing–review and editing. L. Egia-Mendikute: Resources, formal analysis, supervision, validation, investigation, visualization, methodology, writing–review and editing. L. Pérez-Gutiérrez: Resources, formal analysis, validation, investigation, visualization, methodology, writing–review and editing. S.Y. Lee: Resources, formal analysis, validation, investigation, visualization, methodology. A. de Blas: Resources, formal analysis, validation, investigation, visualization, methodology. A. García-del Río: Resources, formal analysis, validation, investigation, visualization, methodology. S.R. Zanetti: Formal analysis, validation, investigation, visualization, methodology. A. Antoñana-Vildosola: Resources, formal analysis, validation, investigation, visualization, methodology. A. Barreira-Manrique: Resources, formal analysis, validation, visualization, methodology. A. Bosch: Resources, Formal analysis, validation, investigation, visualization, methodology. J. Etxaniz-Díaz de Durana: Resources, formal analysis, validation, investigation, visualization, methodology. E. Prieto-Fernández: Resources, data curation, software, formal analysis, validation, investigation, visualization, methodology. M. Serrano-Maciá: Resources, formal analysis, validation, investigation, visualization, methodology. N. Goikoetxea-Usandizaga: Resources, formal analysis, methodology. M. Azkargorta: Resources, data curation, software, formal analysis, methodology. F. Elortza: Resources, data curation, software, formal analysis, methodology. T. Gruber: Resources, formal analysis, validation, investigation, visualization, methodology. S. Peer: Resources, formal analysis, validation, investigation, visualization, methodology. G. Baier: Resources, supervision, funding acquisition. A. Woodhoo: Resources, formal analysis, validation. M.L. Martínez-Chantar: Conceptualization, resources, formal analysis, supervision, validation, investigation, visualization, methodology. A. Palazon: Conceptualization, resources, formal analysis, supervision, funding acquisition, validation, investigation, visualization, methodology, writing–original draft, project administration, writing–review and editing.

This research was supported by the following: Ministerio de Ciencia, Innovación y Cultura (MICIU)/Agencia Española de Investigación (AEI; 10.13039/501100011033) financed “PID2019-107956RA-I00” project (A.P); MICIU/AEI/10.13039/501100011033 and European Union Next Generation EU/PRTR financed “PDC2022-133300-I00” (A. Palazon), “TED2021-129433B-C21” (A. Palazon) and “JDC2022-048612-I” (L. Pérez-Gutiérrez) projects; MICIU/AEI (10.13039/501100011033) and “Fondo Europeo de Desarrollo Regional (FEDER)”/UR financed “PID2022-139344OB-I00” project (A. Palazon); MICIU/AEI/10.13039/501100011033 and “El Fondo Social Europeo (FSE) invierte en tu futuro” financed “RYC2018-024183-I” (A. Palazon) and “PRE2020-092342” (P. Velasco-Beltrán) grants; AEI financed “RYC2021-031213-I” grant (E. Prieto-Fernández), FSE+ financed “PRE2022-104817” (J. Etxaniz-Díaz de Durana) predoctoral grant; Ministerio de Ciencia, Innovación y Universidades MICINN: PID2023-146933OB-100 funded by MCIU /AEI /10.13039/501100011033/FEDER, UE, as part of Plan Estatal de Investigación Científica y Técnica e Innovación (M.L. Martínez-Chantar); MCIN/AEI/10.13039/501100011033 financed the grant CEX2021-001136-S (M.L. Martínez-Chantar); the European Research Council (ERC) under the European Union’s Horizon 2020 research and innovation program financed grant agreement “no 804236” (A. Palazon); the ERC awarded to G. Baier with “ERC-2018-ADG 786462-HOPE” grant; the Provincial Section of Bizkaia and of the Scientific Foundation of the Spanish Association against cancer (AECC) financed “PRDVZ21640DEBL” (A. de Blas) and “PRDVZ1900BOSCH” (A. Bosch) predoctoral grants; the Scientific Foundation of AECC financed “LABAE211744PALA” project (A. Palazon); “La Caixa” Foundation financed the “HR21-00925” (A. Palazon); “La Caixa” Foundation under the program of INPhINIT fellowships financed Doctoral INPhINIT Fellowship – Retaining under the code “LCF/BQ/DR20/11790022” (A. Antoñana-Vildosola); “XVI” BECA FERO (A. Palazon); the Basque Government financed “PRE_2019_1_0320” predoctoral grant (B. Jiménez-Lasheras) through the “Programa Predoctoral de Formación de Personal Investigador No doctor del Departamento de Educación,” “2023333027” project (A. Palazon) through the “Programa Ayudas a proyectos de investigación y desarrollo en salud del Departamento de Salud,” and “AF-W1-2019-00012” through Programa Bikaintek (A. García-del Río).

Note: Supplementary data for this article are available at Cancer Immunology Research Online (http://cancerimmunolres.aacrjournals.org/).

1.
Li
W
,
Li
F
,
Zhang
X
,
Lin
H-K
,
Xu
C
.
Insights into the post-translational modification and its emerging role in shaping the tumor microenvironment
.
Signal Transduct Target Ther
2021
;
6
:
422
.
2.
Shah
K
,
Al-Haidari
A
,
Sun
J
,
Kazi
JU
.
T cell receptor (TCR) signaling in health and disease
.
Signal Transduct Target Ther
2021
;
6
:
412
.
3.
Liu
J
,
Cheng
Y
,
Zheng
M
,
Yuan
B
,
Wang
Z
,
Li
X
, et al
.
Targeting the ubiquitination/deubiquitination process to regulate immune checkpoint pathways
.
Signal Transduct Target Ther
2021
;
6
:
28
.
4.
Friend
SF
,
Deason-Towne
F
,
Peterson
LK
,
Berger
AJ
,
Dragone
LL
.
Regulation of T cell receptor complex-mediated signaling by ubiquitin and ubiquitin-like modifications
.
Am J Clin Exp Immunol
2014
;
3
:
107
23
.
5.
Sun
Y
,
Wei
W
,
Jin
J
.
Cullin-RING ligases and protein neddylation
.
Singapore
:
Springer
;
2020
.
VIII
. p.
372
.
6.
Rabut
G
,
Peter
M
.
Function and regulation of protein neddylation. ‘protein modifications: beyond the usual suspects’ review series
.
EMBO Rep
2008
;
9
:
969
76
.
7.
Zhou
L
,
Jiang
Y
,
Luo
Q
,
Li
L
,
Jia
L
.
Neddylation: a novel modulator of the tumor microenvironment
.
Mol Cancer
2019
;
18
:
77
.
8.
Duda
DM
,
Borg
LA
,
Scott
DC
,
Hunt
HW
,
Hammel
M
,
Schulman
BA
.
Structural insights into NEDD8 activation of cullin-RING ligases: conformational control of conjugation
.
Cell
2008
;
134
:
995
1006
.
9.
Petroski
MD
,
Deshaies
RJ
.
Function and regulation of cullin-RING ubiquitin ligases
.
Nat Rev Mol Cell Biol
2005
;
6
:
9
20
.
10.
Bailly
AP
,
Perrin
A
,
Serrano-Macia
M
,
Maghames
C
,
Leidecker
O
,
Trauchessec
H
, et al
.
The balance between mono- and NEDD8-chains controlled by NEDP1 upon DNA damage is a regulatory module of the HSP70 ATPase activity
.
Cell Rep
2019
;
29
:
212
24.e8
.
11.
Xirodimas
DP
,
Saville
MK
,
Bourdon
J-C
,
Hay
RT
,
Lane
DP
.
Mdm2-mediated NEDD8 conjugation of p53 inhibits its transcriptional activity
.
Cell
2004
;
118
:
83
97
.
12.
Oved
S
,
Mosesson
Y
,
Zwang
Y
,
Santonico
E
,
Shtiegman
K
,
Marmor
MD
, et al
.
Conjugation to Nedd8 instigates ubiquitylation and down-regulation of activated receptor tyrosine kinases
.
J Biol Chem
2006
;
281
:
21640
51
.
13.
Zuo
W
,
Huang
F
,
Chiang
YJ
,
Li
M
,
Du
J
,
Ding
Y
, et al
.
c-Cbl-mediated neddylation antagonizes ubiquitination and degradation of the TGF-beta type II receptor
.
Mol Cell
2013
;
49
:
499
510
.
14.
Kim
S
,
Kwon
M
,
Hwang
Y
,
Yoon
J
,
Park
S
,
Kang
HC
.
Stress-induced NEDDylation promotes cytosolic protein aggregation through HDAC6 in a p62-dependent manner
.
iScience
2021
;
24
:
102146
.
15.
Xie
P
,
Peng
Z
,
Chen
Y
,
Li
H
,
Du
M
,
Tan
Y
, et al
.
Author Correction: neddylation of PTEN regulates its nuclear import and promotes tumor development
.
Cell Res
2021
;
31
:
374
.
16.
Xie
P
,
Zhang
M
,
He
S
,
Lu
K
,
Chen
Y
,
Xing
G
, et al
.
The covalent modifier Nedd8 is critical for the activation of Smurf1 ubiquitin ligase in tumorigenesis
.
Nat Commun
2014
;
5
:
3733
.
17.
Guo
Z
,
Wang
S
,
Xie
Y
,
Han
Y
,
Hu
S
,
Guan
H
, et al
.
HUWE1-dependent DNA-PKcs neddylation modulates its autophosphorylation in DNA damage response
.
Cell Death Dis
2020
;
11
:
400
.
18.
Lu
X
,
Kong
X
,
Wu
H
,
Hao
J
,
Li
S
,
Gu
Z
, et al
.
UBE2M-mediated neddylation of TRIM21 regulates obesity-induced inflammation and metabolic disorders
.
Cell Metab
2023
;
35
:
1390
405.e8
.
19.
Olaizola
P
,
Lee-Law
PY
,
Fernandez-Barrena
MG
,
Alvarez
L
,
Cadamuro
M
,
Azkargorta
M
, et al
.
Targeting NAE1-mediated protein hyper-NEDDylation halts cholangiocarcinogenesis and impacts on tumor-stroma crosstalk in experimental models
.
J Hepatol
2022
;
77
:
177
90
.
20.
Salon
C
,
Brambilla
E
,
Brambilla
C
,
Lantuejoul
S
,
Gazzeri
S
,
Eymin
B
.
Altered pattern of Cul-1 protein expression and neddylation in human lung tumours: relationships with CAND1 and cyclin E protein levels
.
J Pathol
2007
;
213
:
303
10
.
21.
Chairatvit
K
,
Ngamkitidechakul
C
.
Control of cell proliferation via elevated NEDD8 conjugation in oral squamous cell carcinoma
.
Mol Cell Biochem
2007
;
306
:
163
9
.
22.
Soucy
TA
,
Dick
LR
,
Smith
PG
,
Milhollen
MA
,
Brownell
JE
.
The NEDD8 conjugation pathway and its relevance in cancer biology and therapy
.
Genes Cancer
2010
;
1
:
708
16
.
23.
Yu
J
,
Huang
W-L
,
Xu
Q-G
,
Zhang
L
,
Sun
S-H
,
Zhou
W-P
, et al
.
Overactivated neddylation pathway in human hepatocellular carcinoma
.
Cancer Med
2018
;
7
:
3363
72
.
24.
Zhao
Y
,
Morgan
MA
,
Sun
Y
.
Targeting Neddylation pathways to inactivate cullin-RING ligases for anticancer therapy
.
Antioxid Redox Signal
2014
;
21
:
2383
400
.
25.
Brownell
JE
,
Sintchak
MD
,
Gavin
JM
,
Liao
H
,
Bruzzese
FJ
,
Bump
NJ
, et al
.
Substrate-assisted inhibition of ubiquitin-like protein-activating enzymes: the NEDD8 E1 inhibitor MLN4924 forms a NEDD8-AMP mimetic in situ
.
Mol Cell
2010
;
37
:
102
11
.
26.
Shi
C-S
,
Kuo
K-L
,
Lin
W-C
,
Chen
M-S
,
Liu
S-H
,
Liao
S-M
, et al
.
Neddylation inhibitor, MLN4924 suppresses angiogenesis in huvecs and solid cancers: in vitro and in vivo study
.
Am J Cancer Res
2020
;
10
:
953
64
.
27.
Torka
P
,
Thiruvengadam
SK
,
Chen
L
,
Wang
X
,
Chen
C
,
Vuong
D
, et al
.
Pevonedistat, a Nedd8-activating enzyme inhibitor, in combination with ibrutinib in patients with relapsed/refractory B-cell non-Hodgkin lymphoma
.
Blood Cancer J
2023
;
13
:
9
.
28.
Saliba
AN
,
Kaufmann
SH
,
Stein
EM
,
Patel
PA
,
Baer
MR
,
Stock
W
, et al
.
Pevonedistat with azacitidine in older patients with TP53-mutated AML: a phase 2 study with laboratory correlates
.
Blood Adv
2023
;
7
:
2360
3
.
29.
Adès
L
,
Girshova
L
,
Doronin
VA
,
Díez-Campelo
M
,
Valcárcel
D
,
Kambhampati
S
, et al
.
Pevonedistat plus azacitidine vs azacitidine alone in higher-risk MDS/chronic myelomonocytic leukemia or low-blast-percentage AML
.
Blood Adv
2022
;
6
:
5132
45
.
30.
Austin
R
,
Smyth
MJ
,
Lane
SW
.
Harnessing the immune system in acute myeloid leukaemia
.
Crit Rev Oncol Hematol
2016
;
103
:
62
77
.
31.
Khaldoyanidi
S
,
Nagorsen
D
,
Stein
A
,
Ossenkoppele
G
,
Subklewe
M
.
Immune biology of acute myeloid leukemia: implications for immunotherapy
.
J Clin Oncol
2021
;
39
:
419
32
.
32.
Peng
X
,
Zhu
X
,
Di
T
,
Tang
F
,
Guo
X
,
Liu
Y
, et al
.
The yin-yang of immunity: immune dysregulation in myelodysplastic syndrome with different risk stratification
.
Front Immunol
2022
;
13
:
994053
.
33.
Wang
X
,
Chen
C
,
Vuong
D
,
Rodriguez-Rodriguez
S
,
Lam
V
,
Roleder
C
, et al
.
Pharmacologic targeting of Nedd8-activating enzyme reinvigorates T-cell responses in lymphoid neoplasia
.
Leukemia
2023
;
37
:
1324
35
.
34.
Best
S
,
Lam
V
,
Liu
T
,
Bruss
N
,
Kittai
A
,
Danilova
OV
, et al
.
Immunomodulatory effects of pevonedistat, a NEDD8-activating enzyme inhibitor, in chronic lymphocytic leukemia-derived T cells
.
Leukemia
2021
;
35
:
156
68
.
35.
Li
Y
,
Zhou
H
,
Liu
P
,
Lv
D
,
Shi
Y
,
Tang
B
, et al
.
SHP2 deneddylation mediates tumor immunosuppression in colon cancer via the CD47/SIRPα axis
.
J Clin Invest
2023
;
133
:
e162870
.
36.
Cheng
M
,
Hu
S
,
Wang
Z
,
Pei
Y
,
Fan
R
,
Liu
X
, et al
.
Inhibition of neddylation regulates dendritic cell functions via Deptor accumulation driven mTOR inactivation
.
Oncotarget
2016
;
7
:
35643
54
.
37.
Serrano-Maciá
M
,
Lachiondo-Ortega
S
,
Iruzubieta
P
,
Goikoetxea-Usandizaga
N
,
Bosch
A
,
Egia-Mendikute
L
, et al
.
Neddylation tunes peripheral blood mononuclear cells immune response in COVID-19 patients
.
Cell Death Discov
2022
;
8
:
316
.
38.
Kim
K
,
Pröbstel
A-K
,
Baumann
R
,
Dyckow
J
,
Landefeld
J
,
Kogl
E
, et al
.
Cell type-specific transcriptomics identifies neddylation as a novel therapeutic target in multiple sclerosis
.
Brain
2021
;
144
:
450
61
.
39.
Palomo-Irigoyen
M
,
Pérez-Andrés
E
,
Iruarrizaga-Lejarreta
M
,
Barreira-Manrique
A
,
Tamayo-Caro
M
,
Vila-Vecilla
L
, et al
.
HuR/ELAVL1 drives malignant peripheral nerve sheath tumor growth and metastasis
.
J Clin Invest
2020
;
130
:
3848
64
.
40.
Castella
M
,
Boronat
A
,
Martín-Ibáñez
R
,
Rodríguez
V
,
Suñé
G
,
Caballero
M
, et al
.
Development of a novel anti-CD19 chimeric antigen receptor: a paradigm for an affordable CAR T cell production at academic institutions
.
Mol Ther Methods Clin Dev
2018
;
12
:
134
44
.
41.
Dull
T
,
Zufferey
R
,
Kelly
M
,
Mandel
RJ
,
Nguyen
M
,
Trono
D
, et al
.
A third-generation lentivirus vector with a conditional packaging system
.
J Virol
1998
;
72
:
8463
71
.
42.
Vogl
AM
,
Phu
L
,
Becerra
R
,
Giusti
SA
,
Verschueren
E
,
Hinkle
TB
, et al
.
Global site-specific neddylation profiling reveals that NEDDylated cofilin regulates actin dynamics
.
Nat Struct Mol Biol
2020
;
27
:
210
20
.
43.
Scharping
NE
,
Rivadeneira
DB
,
Menk
AV
,
Vignali
PDA
,
Ford
BR
,
Rittenhouse
NL
, et al
.
Mitochondrial stress induced by continuous stimulation under hypoxia rapidly drives T cell exhaustion
.
Nat Immunol
2021
;
22
:
205
15
.
44.
Wiśniewski
JR
,
Zougman
A
,
Nagaraj
N
,
Mann
M
.
Universal sample preparation method for proteome analysis
.
Nat Methods
2009
;
6
:
359
62
.
45.
Prazanowska
KH
,
Lim
SB
.
An integrated single-cell transcriptomic dataset for non-small cell lung cancer
.
Sci Data
2023
;
10
:
167
.
46.
Perez-Riverol
Y
,
Bai
J
,
Bandla
C
,
García-Seisdedos
D
,
Hewapathirana
S
,
Kamatchinathan
S
, et al
.
The PRIDE database resources in 2022: a hub for mass spectrometry-based proteomics evidences
.
Nucleic Acids Res
2022
;
50
:
D543
52
.
47.
Andreatta
M
,
Corria-Osorio
J
,
Müller
S
,
Cubas
R
,
Coukos
G
,
Carmona
SJ
.
Interpretation of T cell states from single-cell transcriptomics data using reference atlases
.
Nat Commun
2021
;
12
:
2965
.
48.
Kim
N
,
Kim
HK
,
Lee
K
,
Hong
Y
,
Cho
JH
,
Choi
JW
, et al
.
Single-cell RNA sequencing demonstrates the molecular and cellular reprogramming of metastatic lung adenocarcinoma
.
Nat Commun
2020
;
11
:
2285
.
49.
Jerby-Arnon
L
,
Shah
P
,
Cuoco
MS
,
Rodman
C
,
Su
M-J
,
Melms
JC
, et al
.
A cancer cell program promotes T cell exclusion and resistance to checkpoint blockade
.
Cell
2018
;
175
:
984
97.e24
.
50.
Ma
L
,
Hernandez
MO
,
Zhao
Y
,
Mehta
M
,
Tran
B
,
Kelly
M
, et al
.
Tumor cell biodiversity drives microenvironmental reprogramming in liver cancer
.
Cancer Cell
2019
;
36
:
418
30.e6
.
51.
Zhang
Y
,
Shi
C-C
,
Zhang
H-P
,
Li
G-Q
,
Li
S-S
.
MLN4924 suppresses neddylation and induces cell cycle arrest, senescence, and apoptosis in human osteosarcoma
.
Oncotarget
2016
;
7
:
45263
74
.
52.
Rosskopf
S
,
Leitner
J
,
Paster
W
,
Morton
LT
,
Hagedoorn
RS
,
Steinberger
P
, et al
.
A Jurkat 76 based triple parameter reporter system to evaluate TCR functions and adoptive T cell strategies
.
Oncotarget
2018
;
9
:
17608
19
.
53.
Battin
C
,
Kaufmann
G
,
Leitner
J
,
Tobias
J
,
Wiedermann
U
,
Rölle
A
, et al
.
NKG2A-checkpoint inhibition and its blockade critically depends on peptides presented by its ligand HLA-E
.
Immunology
2022
;
166
:
507
21
.
54.
Jutz
S
,
Leitner
J
,
Schmetterer
K
,
Doel-Perez
I
,
Majdic
O
,
Grabmeier-Pfistershammer
K
, et al
.
Assessment of costimulation and coinhibition in a triple parameter T cell reporter line: simultaneous measurement of NF-κB, NFAT and AP-1
.
J Immunol Methods
2016
;
430
:
10
20
.
55.
Doi
TS
,
Takahashi
T
,
Taguchi
O
,
Azuma
T
,
Obata
Y
.
NF-kappa B RelA-deficient lymphocytes: normal development of T cells and B cells, impaired production of IgA and IgG1 and reduced proliferative responses
.
J Exp Med
1997
;
185
:
953
61
.
56.
Vaeth
M
,
Feske
S
.
NFAT control of immune function: New Frontiers for an Abiding Trooper
.
F1000Res
2018
;
7
:
260
.
57.
Voisin
A
,
Grinberg-Bleyer
Y
.
The many-sided contributions of NF-κB to T-cell biology in health and disease
.
Int Rev Cell Mol Biol
2021
;
361
:
245
300
.
58.
Meeth
K
,
Wang
JX
,
Micevic
G
,
Damsky
W
,
Bosenberg
MW
.
The YUMM lines: a series of congenic mouse melanoma cell lines with defined genetic alterations
.
Pigment Cell Melanoma Res
2016
;
29
:
590
7
.
59.
Wang
J
,
Perry
CJ
,
Meeth
K
,
Thakral
D
,
Damsky
W
,
Micevic
G
, et al
.
UV-induced somatic mutations elicit a functional T cell response in the YUMMER1.7 mouse melanoma model
.
Pigment Cell Melanoma Res
2017
;
30
:
428
35
.
60.
Kassouf
T
,
Shrivastava
R
,
Meszka
I
,
Bailly
A
,
Polanowska
J
,
Trauchessec
H
, et al
.
Targeting the NEDP1 enzyme to ameliorate ALS phenotypes through stress granule disassembly
.
Sci Adv
2023
;
9
:
eabq7585
.
61.
Lobato-Gil
S
,
Heidelberger
JB
,
Maghames
C
,
Bailly
A
,
Brunello
L
,
Rodriguez
MS
, et al
.
Proteome-wide identification of NEDD8 modification sites reveals distinct proteomes for canonical and atypical NEDDylation
.
Cell Rep
2021
;
34
:
108635
.
62.
Palazon
A
,
Goldrath
AW
,
Nizet
V
,
Johnson
RS
.
HIF transcription factors, inflammation, and immunity
.
Immunity
2014
;
41
:
518
28
.
63.
Curtis
VF
,
Ehrentraut
SF
,
Campbell
EL
,
Glover
LE
,
Bayless
A
,
Kelly
CJ
, et al
.
Stabilization of HIF through inhibition of Cullin-2 neddylation is protective in mucosal inflammatory responses
.
FASEB J
2015
;
29
:
208
15
.
64.
Zhao
Y
,
Xiong
X
,
Jia
L
,
Sun
Y
.
Targeting Cullin-RING ligases by MLN4924 induces autophagy via modulating the HIF1-REDD1-TSC1-mTORC1-DEPTOR axis
.
Cell Death Dis
2012
;
3
:
e386
.
65.
Weidemann
A
,
Johnson
RS
.
Biology of HIF-1alpha
.
Cell Death Differ
2008
;
15
:
621
7
.
66.
The Gene Ontology Consortium
.
The Gene Ontology Resource: 20 years and still GOing strong
.
Nucleic Acids Res
2019
;
47
:
D330
8
.
67.
Cao
J
,
Liao
S
,
Zeng
F
,
Liao
Q
,
Luo
G
,
Zhou
Y
.
Effects of altered glycolysis levels on CD8+ T cell activation and function
.
Cell Death Dis
2023
;
14
:
407
.
68.
Soucy
TA
,
Smith
PG
,
Milhollen
MA
,
Berger
AJ
,
Gavin
JM
,
Adhikari
S
, et al
.
An inhibitor of NEDD8-activating enzyme as a new approach to treat cancer
.
Nature
2009
;
458
:
732
6
.
69.
Zhou
L
,
Lin
X
,
Zhu
J
,
Zhang
L
,
Chen
S
,
Yang
H
, et al
.
NEDD8-conjugating enzyme E2s: critical targets for cancer therapy
.
Cell Death Discov
2023
;
9
:
23
.
70.
Jin
H-S
,
Liao
L
,
Park
Y
,
Liu
Y-C
.
Neddylation pathway regulates T-cell function by targeting an adaptor protein Shc and a protein kinase Erk signaling
.
Proc Natl Acad Sci U S A
2013
;
110
:
624
9
.
71.
Ma
S
,
Ming
Y
,
Wu
J
,
Cui
G
.
Cellular metabolism regulates the differentiation and function of T-cell subsets
.
Cell Mol Immunol
2024
;
21
:
419
35
.
72.
Giannone
G
,
Ghisoni
E
,
Genta
S
,
Scotto
G
,
Tuninetti
V
,
Turinetto
M
, et al
.
Immuno-metabolism and microenvironment in cancer: key players for immunotherapy
.
Int J Mol Sci
2020
;
21
:
4414
.
73.
Dai
M
,
Wang
L
,
Yang
J
,
Chen
J
,
Dou
X
,
Chen
R
, et al
.
LDHA as a regulator of T cell fate and its mechanisms in disease
.
Biomed Pharmacother
2023
;
158
:
114164
.
74.
Gonzalez-Rellan
MJ
,
Fernández
U
,
Parracho
T
,
Novoa
E
,
Fondevila
MF
,
da Silva Lima
N
, et al
.
Neddylation of phosphoenolpyruvate carboxykinase 1 controls glucose metabolism
.
Cell Metab
2023
;
35
:
1630
45.e5
.
75.
Zhou
Q
,
Li
H
,
Li
Y
,
Tan
M
,
Fan
S
,
Cao
C
, et al
.
Inhibiting neddylation modification alters mitochondrial morphology and reprograms energy metabolism in cancer cells
.
JCI Insight
2019
;
4
:
e121582
.
76.
Li
Y
,
Zhang
R
,
Hei
H
.
Advances in post-translational modifications of proteins and cancer immunotherapy
.
Front Immunol
2023
;
14
:
1229397
.
77.
Li
W
,
Qiu
S
,
Chen
J
,
Jiang
S
,
Chen
W
,
Jiang
J
, et al
.
Chimeric antigen receptor designed to prevent ubiquitination and downregulation showed durable antitumor efficacy
.
Immunity
2020
;
53
:
456
70.e6
.
78.
Harris
MJ
,
Chen
H
,
Cai
T
,
Yi
Y
,
Deng
Q
,
Yao
Y
, et al
.
Generation of allogeneic CAR T cells through specific degradation of the T cell antigen receptor by E3 ubiquitin ligase fusion proteins
.
ACS Synth Biol
2022
;
11
:
2029
35
.
79.
Liao
X
,
Li
W
,
Zhou
H
,
Rajendran
BK
,
Li
A
,
Ren
J
, et al
.
The CUL5 E3 ligase complex negatively regulates central signaling pathways in CD8+ T cells
.
Nat Commun
2024
;
15
:
603
.
This open access article is distributed under the Creative Commons Attribution-NonCommercial-NoDerivatives 4.0 International (CC BY-NC-ND 4.0) license.