Constitutive activation of NF-κB is a hallmark of the activated B cell–like (ABC) subtype of diffuse large B-cell lymphoma (DLBCL), owing to upstream signals from the B-cell receptor (BCR) and MYD88 pathways. The linear polyubiquitin chain assembly complex (LUBAC) attaches linear polyubiquitin chains to IκB kinase-γ, a necessary event in some pathways that engage NF-κB. Two germline polymorphisms affecting the LUBAC subunit RNF31 are rare among healthy individuals (∼1%) but enriched in ABC DLBCL (7.8%). These polymorphisms alter RNF31 α-helices that mediate binding to the LUBAC subunit RBCK1, thereby increasing RNF31–RBCK1 association, LUBAC enzymatic activity, and NF-κB engagement. In the BCR pathway, LUBAC associates with the CARD11–MALT1–BCL10 adapter complex and is required for ABC DLBCL viability. A stapled RNF31 α-helical peptide based on the ABC DLBCL–associated Q622L polymorphism inhibited RNF31–RBCK1 binding, decreased NF-κB activation, and killed ABC DLBCL cells, credentialing this protein–protein interface as a therapeutic target.

Significance: We provide genetic, biochemical, and functional evidence that the LUBAC ubiquitin ligase is a therapeutic target in ABC DLBCL, the DLBCL subtype that is most refractory to current therapy. More generally, our findings highlight the role of rare germline-encoded protein variants in cancer pathogenesis. Cancer Discov; 4(4); 480–93. ©2014 AACR.

See related commentary by Grumati and Dikic, p. 394

This article is highlighted in the In This Issue feature, p. 377

Diffuse large B-cell lymphoma (DLBCL) can be divided into two main molecular subtypes, termed activated B cell–like (ABC) and germinal center B cell–like (GCB) DLBCL, which differ in their gene expression profiles, oncogenic abnormalities, and clinical behavior (1, 2). In ABC DLBCL, regulatory pathways normally associated with B-cell activation are constitutively engaged (1). In particular, the NF-κB pathway plays an essential role in its pathogenesis by promoting malignant cell survival and inducing expression of the master regulatory transcription factor IRF4 (3, 4).

Recent genomic and functional studies have elucidated the molecular mechanisms underlying constitutive NF-κB activity in ABC DLBCL, highlighting the central role of the B-cell receptor (BCR) and MYD88 signaling pathways. The involvement of BCR signaling in ABC DLBCL was first revealed by the dependence of these lymphomas on the adapter protein CARD11 (5). In response to BCR signaling, CARD11 forms a multiprotein “CBM” complex with MALT1 and BCL10 and activates the inhibitor of IκB kinase (IKK), thereby triggering the canonical NF-κB pathway. In 10% of ABC DLBCL tumors, CARD11 sustains oncogenic somatic mutations that constitutively activate IKK and NF-κB (6). In other ABC DLBCLs with wild-type (WT) CARD11, CARD11 is, nonetheless, essential for survival, revealing the dependence of these lymphomas on BCR signaling, a phenomenon dubbed “chronic active” BCR signaling (7). In more than 20% of ABC DLBCL cases, mutations in the immunoreceptor tyrosine-based activation motifs (ITAM) of the BCR subunits CD79B and CD79A augment chronic active BCR signaling (7), providing genetic evidence that BCR signaling is central to the pathogenesis of this lymphoma subtype. A second pathway activating NF-κB in ABC DLBCL is mediated by MYD88, the central adapter in Toll-like receptor (TLR) signaling (8). MYD88 silencing is lethal to ABC DLBCL cells due to inhibition of NF-κB and autocrine interleukin (IL)-6/IL-10 signaling through Janus-activated kinase (JAK) and STAT3 (8, 9). In 39% of ABC DLBCL cases, this pathway is activated by somatic, gain-of-function MYD88 mutations (8). The most common MYD88 mutant, L265P, spontaneously coordinates a signaling complex in which IRAK4 phosphorylates IRAK1, leading to IKK and NF-κB activation (8).

Protein ubiquitination is involved in various steps of the NF-κB pathway (10). A recently identified type of polyubiquitin, the linear polyubiquitin chain, plays important roles in NF-κB activation (11–14). This polyubiquitin chain is generated by linkages between the C- and N-terminal amino acids of ubiquitin modules, resulting in a head-to-tail linear polyubiquitin polymer. The E3 ligase complex responsible for linear polyubiquitin chain formation is the linear ubiquitin chain assembly complex (LUBAC), composed of RNF31 (HOIP), RBCK1 (HOIL-1L), and SHARPIN. In the canonical NF-κB pathway, LUBAC specifically recognizes and conjugates linear polyubiquitin chains onto the IKK-γ/NEMO subunit, which is considered to be an essential event that activates IKK and NF-κB (11, 15). Cells derived from both cpdm (Sharpin-mutant) and Rbck1−/− mice have reduced activation of the canonical NF-κB pathway in response to multiple stimuli as well as increased TNF-α–induced apoptosis, highlighting the critical role of LUBAC in NF-κB activation (11–14). Although the full physiologic function of LUBAC is still largely unknown, it seems to regulate B-cell function and innate immune responses (12, 13, 16–19). The fact that the BCR and MYD88 signaling pathways are recurrently targeted by genetic changes in ABC DLBCL suggests that the constitutive activation of NF-κB in this malignancy could depend on LUBAC function.

Although NF-κB is an attractive therapeutic target in ABC DLBCL, no IKK inhibitors have been developed for clinical use due to concerns about the pleiotropic effects of IKK and potential on-target toxicities. Unlike mice with disruption of the genes encoding IKK-β or NEMO, which succumb to massive liver cell death, knockout animals for the LUBAC component RBCK1 are born healthy, suggesting that therapies targeting this pathway might have tolerable side effects. Because ABC DLBCL is the subtype of DLBCL that is most refractory to current therapy (20), new therapeutic strategies are needed. In the present report, we investigate the role of LUBAC in ABC DLBCL and its potential as a therapeutic target.

Enrichment of Two Rare SNPs among ABC DLBCL Tumors

Given the importance of LUBAC activity in NF-κB signaling, we searched for mutations affecting LUBAC components using RNA-seq data from ABC DLBCL biopsies and identified two recurrent missense RNF31 mutations that change glutamine 584 to histidine (Q584H; n = 2) and glutamine 622 to leucine (Q622L; n = 3). Both of these mutations have been identified previously as rare single-nucleotide polymorphisms (SNP); among the healthy individuals studied in the 1000 Genomes Project (n = 1,094; ref. 21) and the Grand Opportunity (GO) Exome Sequencing Project (n = 8,413; ref. 22), Q584H (SNP accession rs184184005) had a minor allele frequency (MAF) of 0.19% and 0.13%, respectively, whereas Q622L (SNP accession rs149481717) had a MAF of 0.24% and 0.49%, respectively. Both SNPs are located in a highly conserved region of RNF31 that encodes the ubiquitin-associated (UBA) domain, which interacts with the ubiquitin-like (UBL) domain of RBCK1, leading to LUBAC enzyme formation (Fig. 1A and B; refs. 11, 23).

Figure 1.

Enrichment of two rare SNPs among ABC DLBCL tumors. A, amino acid sequence (based on accession NP_060469) of a region of the UBA domain of RNF31 showing the residues altered by two SNPs, Q584H and Q622L, and the number and type of lymphoma biopsies in which they were identified. B, location of the residues altered by the RNF31 SNPs in two views of the three-dimensional structure of the RNF31 UBA domain. C, frequencies of RNF31 SNPs in biopsy samples from different lymphoma subtypes.

Figure 1.

Enrichment of two rare SNPs among ABC DLBCL tumors. A, amino acid sequence (based on accession NP_060469) of a region of the UBA domain of RNF31 showing the residues altered by two SNPs, Q584H and Q622L, and the number and type of lymphoma biopsies in which they were identified. B, location of the residues altered by the RNF31 SNPs in two views of the three-dimensional structure of the RNF31 UBA domain. C, frequencies of RNF31 SNPs in biopsy samples from different lymphoma subtypes.

Close modal

We resequenced RNF31 exon 10, which includes these SNPs, in 561 biopsy samples of various lymphoma subtypes. In 103 ABC DLBCL biopsies, we detected Q584H in 2 cases (1.94%) and Q622L in 6 cases (5.83%), with an overall frequency of 7.77% [95% confidence interval (CI), 3.41%–14.73%], which is 8.22-fold (95% CI, 4.05–16.65) higher than in healthy individuals studied in the GO Exome Sequencing Project (22) and the 1000 Genomes Project (P = 1.02E−5; Fig. 1C; Supplementary Table S1; ref. 21). Among 458 samples of other lymphoid malignancies, 3 Q622L cases were identified, one each in the GCB subtype of DLBCL, follicular lymphoma, and Hodgkin lymphoma, and no cases had Q584H, yielding a frequency of both SNPs in non–ABC DLBCL cases of 0.66%, which is similar to the frequency in the healthy cohorts (0.95%), but 11.9-fold lower than the frequency in ABC DLBCL (P = 1.03E−4). Of note, all of the ABC DLBCL cases carrying these SNPs had either a MYD88 mutation or genetic aberrations affecting A20, whereas only one had a mutation in the CD79B subunit of the BCR (Supplementary Table S2). In 5 ABC DLBCL cases with available germline DNA, both Q584H and Q622L were confirmed to be germline variants (Supplementary Fig. S1).

LUBAC Is Essential for NF-κB Activity in ABC DLBCL

RNA interference–mediated depletion of RNF31 and SHARPIN was toxic for most of ABC DLBCL lines, but had little effect on the GCB DLBCL lines tested (Fig. 2A). The attachment of linear ubiquitin chains to the IKK-γ/NEMO subunit is required for NF-κB activity in response to various stimuli (11, 15). In ABC DLBCL cells, NEMO was constitutively modified by polyubiquitin, and depletion of RNF31 or SHARPIN decreased this modification (Fig. 2B). Accordingly, depletion of these LUBAC components decreased two indicators of IKK activity, phosphorylation of IKK-β and its substrate IκBα (Fig. 2B). The activity of IKK-β can be measured using a reporter construct in which IκBα is fused to luciferase (24). Knockdown of RNF31 in an ABC DLBCL line caused a rise in the IκBα luciferase reporter, indicating IKK-β inhibition (Fig. 2C). Likewise, RNF31 and SHARPIN depletion decreased nuclear NF-κB p65 DNA binding (Fig. 2D), and RNF31 depletion reduced NF-κB transcriptional activity in ABC DLBCL, as indicated by a luciferase reporter driven by an NF-κB response element (Fig. 2E). Hence, LUBAC is essential for maintaining NF-κB activity and viability of ABC DLBCL cells.

Figure 2.

LUBAC is essential for NF-κB activity in ABC DLBCL. A, RNF31 and SHARPIN shRNAs are selectively toxic for ABC DLBCL lines. Shown is the fraction of viable GFP+, shRNA-expressing cells relative to the total live cell fraction at the indicated times following shRNA induction, normalized to the day 0 values. B, lysates were prepared using 1% SDS from HBL1 ABC DLBCL cells expressing the indicated shRNAs, diluted, and then subjected to anti-NEMO immunoprecipitation, followed by immunoblotting for indicated proteins (left). Relative NEMO ubiquitination signal intensity was determined by densitometric analysis (middle). Lysates were additionally analyzed by immunoblotting for the indicated proteins (right). C, relative IκBα luciferase reporter activity in TMD8 ABC DLBCL cells induced for various days to express the indicated shRNAs. A specific IKK-β inhibitor (MLN120B; 10 μmol/L) was included as a positive control. D, nuclear fractions prepared from HBL1 cells expressing the indicated shRNAs were assayed for NF-κB p65 DNA-binding activity by ELISA. E, relative activity of an NF-κB–dependent luciferase reporter in the ABC DLBCL lines expressing the indicated shRNAs. All error bars, SEM of triplicates.

Figure 2.

LUBAC is essential for NF-κB activity in ABC DLBCL. A, RNF31 and SHARPIN shRNAs are selectively toxic for ABC DLBCL lines. Shown is the fraction of viable GFP+, shRNA-expressing cells relative to the total live cell fraction at the indicated times following shRNA induction, normalized to the day 0 values. B, lysates were prepared using 1% SDS from HBL1 ABC DLBCL cells expressing the indicated shRNAs, diluted, and then subjected to anti-NEMO immunoprecipitation, followed by immunoblotting for indicated proteins (left). Relative NEMO ubiquitination signal intensity was determined by densitometric analysis (middle). Lysates were additionally analyzed by immunoblotting for the indicated proteins (right). C, relative IκBα luciferase reporter activity in TMD8 ABC DLBCL cells induced for various days to express the indicated shRNAs. A specific IKK-β inhibitor (MLN120B; 10 μmol/L) was included as a positive control. D, nuclear fractions prepared from HBL1 cells expressing the indicated shRNAs were assayed for NF-κB p65 DNA-binding activity by ELISA. E, relative activity of an NF-κB–dependent luciferase reporter in the ABC DLBCL lines expressing the indicated shRNAs. All error bars, SEM of triplicates.

Close modal

Role of LUBAC in CBM Complex–Mediated NF-κB Activation

We next investigated the role of LUBAC in the BCR and MYD88 pathways, which govern NF-κB activity in ABC DLBCL. By coimmunoprecipitation, RNF31 associated with MALT1 and, to a lesser extent, IRAK1 in ABC DLBCL lines, suggesting that the LUBAC complex could play a role in both pathways (Fig. 3A). Using an antibody specific for linear ubiquitin (11, 13), this modification was detectable on IKK-γ/NEMO immunoprecipitated from ABC DLBCL cells, as expected, but also on IRAK1 (Supplementary Fig. S2). Neither protein was modified by linear ubiquitin in the control GCB DLBCL line. In contrast, linear ubiquitin was not detectable in immunoprecipitates of MALT1 or CARD11.

Figure 3.

LUBAC is involved in CBM complex mediated NF-κB activation in ABC DLBCL. A, ABC DLBCL lines engineered to express Myc epitope-tagged RNF31 were immunoprecipitated using antibodies to IRAK1 or MALT1, or control immunoglobulin G (IgG; ctrl.) and were analyzed by immunoblotting for the indicated proteins. B, HBL1 cells were engineered to express the indicated shRNAs, treated with MALT1 inhibitor Z-VRPR-fmk (75 μmol/L) for 24 hours or left untreated. Whole-cell lysate was immunoblotted for indicated proteins. MALT1-dependent A20 cleavage products are indicated. C, HBL1 cells expressing the indicated shRNAs were activated by anti-IgM treatment (10 μg/mL) for indicated times and analyzed by immunoblotting for the indicated proteins. D, viability of ABC DLBCL lines expressing control or RNF31 shRNAs were treated with DMSO, ibrutinib (1 nmol/L), or lenalidomide (2 μmol/L) and analyzed by fluorescence-activated cell sorting (FACS) for viable GFP+/shRNA-expressing cells over a time course.

Figure 3.

LUBAC is involved in CBM complex mediated NF-κB activation in ABC DLBCL. A, ABC DLBCL lines engineered to express Myc epitope-tagged RNF31 were immunoprecipitated using antibodies to IRAK1 or MALT1, or control immunoglobulin G (IgG; ctrl.) and were analyzed by immunoblotting for the indicated proteins. B, HBL1 cells were engineered to express the indicated shRNAs, treated with MALT1 inhibitor Z-VRPR-fmk (75 μmol/L) for 24 hours or left untreated. Whole-cell lysate was immunoblotted for indicated proteins. MALT1-dependent A20 cleavage products are indicated. C, HBL1 cells expressing the indicated shRNAs were activated by anti-IgM treatment (10 μg/mL) for indicated times and analyzed by immunoblotting for the indicated proteins. D, viability of ABC DLBCL lines expressing control or RNF31 shRNAs were treated with DMSO, ibrutinib (1 nmol/L), or lenalidomide (2 μmol/L) and analyzed by fluorescence-activated cell sorting (FACS) for viable GFP+/shRNA-expressing cells over a time course.

Close modal

Chronic active BCR signaling in ABC DLBCL causes MALT1 to proteolytically cleave A20, a negative regulator of NF-κB signaling (25, 26). Knockdown of RNF31 decreased A20 proteolysis in ABC DLBCL lines, implicating LUBAC in this regulatory process (Fig. 3B). Acute BCR cross-linking by anti-immunoglobulin M (IgM) antibodies in a GCB DLBCL line (BJAB) or in an ABC DLBCL line (HBL1) rapidly increased IKK-β phosphorylation, but knockdown of RNF31 compromised this induction, reinforcing the view that LUBAC plays a key role in NF-κB activation during BCR signaling (Fig. 3C and Supplementary Fig. S3A). In keeping with these results, ABC DLBCL lines depleted of RNF31 were sensitized to the Bruton agammaglobulinemia tyrosine kinase (BTK) kinase inhibitor ibrutinib, which blocks chronic active BCR signaling in ABC DLBCL (7), and to lenalidomide, which reduces CARD11 levels by inhibiting IRF4 (Fig. 3D; Supplementary Fig. S3B; ref. 4). Together, these results show that LUBAC is associated with the CBM complex and contributes to BCR signaling in ABC DLBCL cells.

RNF31 SNPs Promote NF-κB Activity in ABC DLBCL

A small region of the RNF31 UBA domain, from amino acids 579 to 623, binds to the UBL domain of RBCK1 (23). The RNF31 Q584H and Q622L mutants reside in this region, suggesting that they might promote LUBAC complex formation and subsequent NF-κB activation. When these RNF31 mutants or WT RNF31 were expressed in ABC DLBCL cells at equivalent levels, Q622L and Q584H increased the activity of an NF-κB–driven luciferase reporter more effectively than WT RNF31 (Fig. 4A). Expression levels of two well-known NF-κB target genes, NFKBIA and IRF4, were elevated by the RNF31 mutants more than by WT RNF31 (Fig. 4B). The RNF31 mutants were also more active in stimulating IKK activity than WT RNF31, as judged by the IκBα luciferase reporter (Fig. 4C), and, accordingly, were also superior in stimulating phosphorylation of IKK-β and its substrate IκBα (Fig. 4D) and in inducing nuclear NF-κB p65 DNA-binding activity (Fig. 4E). When expressed in the GCB DLBCL BJAB, the RNF31 mutants induced expression of the NF-κB target CD83, especially in response to anti–IgM-induced BCR activation (Fig. 4F), supporting the notion that LUBAC contributes to BCR-induced engagement of NF-κB (Fig. 3C and Supplementary Fig. S3A).

Figure 4.

RNF31 SNPs promote NF-κB activity in ABC DLBCL. A, NF-κB–driven luciferase reporter activity in HBL1 and TMD8 lines engineered to express the indicated Myc epitope-tagged RNF31 isoforms. B, control or RNF31 shRNAs were inducibly expressed in HBL1 cells that had been transduced with rescue vectors expressing RNF31 isoforms or with an empty vector. Indicated mRNA expression was quantified by quantitative PCR (qPCR) and normalized to β2-microglobulin mRNA levels. C, relative IκBα luciferase reporter activity was measured in TMD8 cells induced to express the indicated Myc-tagged RNF31 isoforms. D, cells prepared as in B were analyzed by immunoblotting for the indicated proteins. E, nuclear NF-κB p65 DNA-binding activity was determined by ELISA in HBL1 cells selected for expression of the indicated Myc epitope-tagged RNF31 isoforms. F, flow cytometry histograms of the expression of the NF-κB target gene CD83 in BJAB cells expressing vector only or the indicated exogenous RNF31 isoforms, with and without anti-IgM stimulation. Right, summary of CD83 expression in BJAB cells expressing vector only or RNF31 isoforms, either in unstimulated cells or anti–IgM-stimulated cells. G, HBL1 cells engineered to express indicated Myc epitope-tagged RNF31 isoforms were analyzed by immunoblotting for the indicated proteins (top). The relative A20 cleavage signal intensity was determined by densitometric analysis (bottom). H, cells prepared as in G were subjected to anti-MALT1 immunoprecipitation. Immunoprecipitated proteins or whole-cell lysates were analyzed by immunoblotting for the indicated proteins. All error bars, SEM of triplicates.

Figure 4.

RNF31 SNPs promote NF-κB activity in ABC DLBCL. A, NF-κB–driven luciferase reporter activity in HBL1 and TMD8 lines engineered to express the indicated Myc epitope-tagged RNF31 isoforms. B, control or RNF31 shRNAs were inducibly expressed in HBL1 cells that had been transduced with rescue vectors expressing RNF31 isoforms or with an empty vector. Indicated mRNA expression was quantified by quantitative PCR (qPCR) and normalized to β2-microglobulin mRNA levels. C, relative IκBα luciferase reporter activity was measured in TMD8 cells induced to express the indicated Myc-tagged RNF31 isoforms. D, cells prepared as in B were analyzed by immunoblotting for the indicated proteins. E, nuclear NF-κB p65 DNA-binding activity was determined by ELISA in HBL1 cells selected for expression of the indicated Myc epitope-tagged RNF31 isoforms. F, flow cytometry histograms of the expression of the NF-κB target gene CD83 in BJAB cells expressing vector only or the indicated exogenous RNF31 isoforms, with and without anti-IgM stimulation. Right, summary of CD83 expression in BJAB cells expressing vector only or RNF31 isoforms, either in unstimulated cells or anti–IgM-stimulated cells. G, HBL1 cells engineered to express indicated Myc epitope-tagged RNF31 isoforms were analyzed by immunoblotting for the indicated proteins (top). The relative A20 cleavage signal intensity was determined by densitometric analysis (bottom). H, cells prepared as in G were subjected to anti-MALT1 immunoprecipitation. Immunoprecipitated proteins or whole-cell lysates were analyzed by immunoblotting for the indicated proteins. All error bars, SEM of triplicates.

Close modal

In keeping with this hypothesis, RNF31 mutants were more effective than WT RNF31 in stimulating MALT1-dependent cleavage of A20 in ABC DLBCL cells (Fig. 4G). Although both mutant and WT RNF31 isoforms interacted with MALT1 equivalently, A20 was more effectively recruited to the CBM complex in ABC DLBCL cells expressing mutant RNF31 than in cells expressing WT RNF31 (Fig. 4H). Recent reports demonstrated that zinc finger 7 of A20 binds linear polyubiquitin chains preferentially, thereby facilitating its recruitment to receptor signaling complexes containing LUBAC and IKK (27, 28). In light of this, our results suggested that the RNF31 mutants may promote A20 cleavage by stimulating LUBAC ubiquitination activity and increasing A20 recruitment to the CBM complex.

RNF31 SNPs Enhance LUBAC Formation and E3 Ligase Activity

To investigate the possibility that the RNF31 mutants have elevated E3 ligase activity, we studied LUBAC-mediated ubiquitination of IKK-γ/NEMO, a modification that is necessary for IKK activation (11). RNF31 depletion decreased the constitutive NEMO linear ubiquitination in the HBL1 ABC DLBCL line, and the RNF31 mutants were more effective in restoring NEMO linear ubiquitination than WT RNF31 (Fig. 5A). We next transduced HBL1 ABC DLBCL cells with Myc epitope-tagged WT or mutant RNF31 isoforms and evaluated E3 ligase activity in anti-Myc immunoprecipitates using an ELISA-based assay as well as in an in vitro polyubiquitin chain formation assay. Although Myc-tagged WT RNF31 immunoprecipitates had ubiquitin ligase activity, LUBAC complexes formed with Myc-tagged RNF31 mutants were more active (Fig. 5B and C). The enzyme activity of LUBAC relies largely on the interaction between RNF31 and RBCK1 (29, 30), suggesting that the RNF31 mutants might increase LUBAC formation. By coimmunoprecipitation, the RNF31 Q662L and Q584H interacted 2.07-fold and 1.4-fold more effectively with RBCK1, respectively, than did WT RNF31 (P < 0.0001; Fig. 5D and E).

Figure 5.

Gain-of-function conferred by RNF31 SNPs. A, control or RNF31 shRNAs were inducibly expressed in HBL1 cells that had been transduced with rescue vectors expressing RNF31 isoforms or with an empty vector. Doxycycline (Dox)-induced cells were lysed in 1% SDS, diluted, and then subjected to immunoprecipitation with an anti-NEMO antibody, followed by immunoblotting for indicated proteins (top). The relative NEMO linear ubiquitination signal intensity was determined by densitometric analysis (bottom). Also shown are immunoblots for the indicated proteins in whole-cell lysates from the same cells. B, HBL1 cells engineered to express indicated Myc epitope-tagged RNF31 isoforms were subject to anti-Myc immunoprecipitation followed by elution with Myc peptides. The elutions were examined in an E3 ligase activity ELISA assay (top) or by immunoblotting for the indicated proteins. C, protein prepared as in B were used in an in vitro ubiqutination assay, the products of which were analyzed by immunoblotting for the indicated proteins. D, anti-Myc immunoprecipitates prepared as in B or whole-cell lysates were analyzed by immunoblotting for the indicated proteins. E, densitometic quantititation of coimmunoprecipitation experiments demonstrating the association of RNF31 isoforms and RBCK as in D. All error bars, SEM of replicate experiments [n = 3 for all experiments except E (n = 11)].

Figure 5.

Gain-of-function conferred by RNF31 SNPs. A, control or RNF31 shRNAs were inducibly expressed in HBL1 cells that had been transduced with rescue vectors expressing RNF31 isoforms or with an empty vector. Doxycycline (Dox)-induced cells were lysed in 1% SDS, diluted, and then subjected to immunoprecipitation with an anti-NEMO antibody, followed by immunoblotting for indicated proteins (top). The relative NEMO linear ubiquitination signal intensity was determined by densitometric analysis (bottom). Also shown are immunoblots for the indicated proteins in whole-cell lysates from the same cells. B, HBL1 cells engineered to express indicated Myc epitope-tagged RNF31 isoforms were subject to anti-Myc immunoprecipitation followed by elution with Myc peptides. The elutions were examined in an E3 ligase activity ELISA assay (top) or by immunoblotting for the indicated proteins. C, protein prepared as in B were used in an in vitro ubiqutination assay, the products of which were analyzed by immunoblotting for the indicated proteins. D, anti-Myc immunoprecipitates prepared as in B or whole-cell lysates were analyzed by immunoblotting for the indicated proteins. E, densitometic quantititation of coimmunoprecipitation experiments demonstrating the association of RNF31 isoforms and RBCK as in D. All error bars, SEM of replicate experiments [n = 3 for all experiments except E (n = 11)].

Close modal

Targeting the RBCK1–RNF31 Interface with Stapled α-Helical RNF31 Peptides

We next considered the possibility that the RBCK1–RNF31 interaction surface that is altered by the RNF31 polymorphisms could be a therapeutic target. To address this, we synthesized a series of peptides modeled on RNF31 α-helices 8 and 9 (Fig. 1A and B) that reside at the RBCK1–RNF31 interface using “hydrocarbon stapling” (31) to stabilize their α-helical structure (Fig. 6A). Stapling of the amino-terminal α-helix 8 in the RNF31 WT and Q622L peptides increased their α-helical character, as expected (Fig. 6B). In cultures of HBL1 cells exposed to fluorescein isothiocyanate (FITC)–conjugated derivatives of these peptides, all cells internalized these peptides to roughly equivalent levels, as judged by flow cytometry and confocal microscopy (Supplementary Fig. S4A and S4B). The ability of these FITC-conjugated peptides to bind to recombinant RBCK1 in vitro was examined quantitatively using fluorescence polarization spectroscopy. Although both stapled peptides interacted more strongly with RBCK1 than did the unstapled peptide, the stapled FITC–RNF31 N-Q622L peptide was clearly superior to the stapled WT RNF31 peptide (Fig. 6C). We next set up a competition assay in which binding of FITC–RNF31 N-Q622L to RBCK1 was inhibited by increasing concentrations of unlabeled peptides. On the basis of the difference in IC50 values, the N-Q622L RNF31 peptide had an approximately 6-fold higher relative affinity for RBCK1 than the WT RNF31 peptide (Fig. 6D), in keeping with the previous coimmunoprecipitation results (Fig. 5D).

Figure 6.

Targeting the RBCK1–RNF31 interface using stapled α-helical RNF31 peptides. A, schematic and sequences of RNF31 stapled peptides. The asterisks indicate the location of the hydrocarbon cross-linker. B, circular dichroism spectroscopy of RNF31 peptides. Dips in the curves at 205 and 225 nm are indicative of α-helical structure. C, fluorescence polarization assay of FITC–labeled RNF31 peptides (14.1 nmol/L) binding to recombinant RBCK1. D, competitive fluorescence polarization assay in which binding FITC–RNF31 N-Q622L (14.1 nmol/L) to RBCK1 (0.7 μmol/L) was inhibited by the indicated concentrations of unlabeled peptides.

Figure 6.

Targeting the RBCK1–RNF31 interface using stapled α-helical RNF31 peptides. A, schematic and sequences of RNF31 stapled peptides. The asterisks indicate the location of the hydrocarbon cross-linker. B, circular dichroism spectroscopy of RNF31 peptides. Dips in the curves at 205 and 225 nm are indicative of α-helical structure. C, fluorescence polarization assay of FITC–labeled RNF31 peptides (14.1 nmol/L) binding to recombinant RBCK1. D, competitive fluorescence polarization assay in which binding FITC–RNF31 N-Q622L (14.1 nmol/L) to RBCK1 (0.7 μmol/L) was inhibited by the indicated concentrations of unlabeled peptides.

Close modal

Biologic Effects of RNF31 Stapled Peptides in ABC DLBCL

The N-Q622L peptide prevented endogenous LUBAC formation in ABC DLBCL cells to a greater degree than WT RNF31, as judged by RNF31 coimmunoprecipitation with RBCK1, and the unstapled peptide had little, if any, effect (Supplementary Fig. S5A). NF-κB pathway activity in ABC DLBCL cells, as measured by the IκBα luciferase assay for IKK-β activity and the NF-κB–driven luciferase reporter, was inhibited by the stapled peptides, with the N-Q622L being more active than N-WT, whereas the unstapled peptide and an unrelated stapled peptide were inactive (Fig. 7A and B). RNF31 N-Q622L killed two ABC DLBCL lines in a dose-dependent fashion but had no toxicity for two GCB DLBCL lines (Fig. 7C). Ectopic expression of a constitutively active IKK-β mutant mitigated the effects of RNF31 N-Q622L on ABC DLBCL viability, consistent with IKK being a major target of LUBAC activity in this lymphoma subtype (Supplementary Fig. S5B). Moreover, this stapled peptide sensitized ABC DLBCL lines to the lethal effects of the BTK inhibitor ibrutinib (Fig. 7C). Besides promoting the survival of ABC DLBCL cells, the NF-κB pathway is well known to inhibit the cytotoxic action of chemotherapy (32). Furthermore, NF-κB is activated by chemotherapy-induced DNA damage (33), and LUBAC activity is essential for this stress response (34). In keeping with these reports, depletion of RNF31 in ABC DLBCL cells impaired NF-κB activation in response to the topoisomerase inhibitor etoposide, as measured by phosphorylation of IKK-β and IκBα, whereas ectopic provision of either of the RNF31 mutants restored the NF-κB response to a greater extent than WT RNF31 (Fig. 7D). The RNF31 stapled peptides cooperated with etoposide in killing ABC DLBCL cells, with the N-Q622L stapled peptide having more activity than the WT version (Fig. 7E).

Figure 7.

Biologic effects of RNF31 stapled peptides in ABC DLBCL. A, NF-κB–dependent luciferase reporter activity in HBL1 cells treated for 2 days with DMSO or the indicated peptides. Data are normalized to DMSO-treated cells. B, IκBα luciferase reporter activity in TMD8 cells treated for 3 days with DMSO or the indicated peptides. Data are normalized to DMSO-treated cells. C, viability of DLBCL lines treated with stapled RNF31 Q622L peptide at the indicated concentrations ± ibrutinib (0.5 nmol/L) for 3 days, normalized to DMSO-treated cells. D, HBL1 cells prepared as in Fig. 4B were exposed to etoposide (5 μmol/L) or DMSO for 1.5 hours. Whole-cell lysates were blotted for the indicated proteins. E, viability of TMD8 cells treated for 4 days with DMSO or the indicated peptides at various concentrations ± etoposide (100 nmol/L), normalized to viability of DMSO-treated cells. Neg. ctrl., negative control stapled peptide. All error bars denote SEM of triplicates.

Figure 7.

Biologic effects of RNF31 stapled peptides in ABC DLBCL. A, NF-κB–dependent luciferase reporter activity in HBL1 cells treated for 2 days with DMSO or the indicated peptides. Data are normalized to DMSO-treated cells. B, IκBα luciferase reporter activity in TMD8 cells treated for 3 days with DMSO or the indicated peptides. Data are normalized to DMSO-treated cells. C, viability of DLBCL lines treated with stapled RNF31 Q622L peptide at the indicated concentrations ± ibrutinib (0.5 nmol/L) for 3 days, normalized to DMSO-treated cells. D, HBL1 cells prepared as in Fig. 4B were exposed to etoposide (5 μmol/L) or DMSO for 1.5 hours. Whole-cell lysates were blotted for the indicated proteins. E, viability of TMD8 cells treated for 4 days with DMSO or the indicated peptides at various concentrations ± etoposide (100 nmol/L), normalized to viability of DMSO-treated cells. Neg. ctrl., negative control stapled peptide. All error bars denote SEM of triplicates.

Close modal

We report two rare germline polymorphisms affecting the LUBAC subunit RNF31 that were enriched among patients with ABC DLBCL relative to patients with other lymphoma subtypes and to healthy individuals. This genetic observation uncovered an essential role for LUBAC enzyme function in maintaining constitutive NF-κB activity in ABC DLBCL cells, which is the central feature of its pathogenesis. The ABC DLBCL–associated SNPs, which alter the domain of RNF31 that interacts with RBCK1, enhance LUBAC complex formation, ubiquitin ligase activity, and stimulation of the NF-κB pathway. We credentialed the RNF31–RBCK1 interface as a therapeutic target using stapled α-helical peptides based on the RNF31 SNPs. This work highlights the potential for rare SNPs in the human population to play a pathogenic role in human disease.

The ABC DLBCL–associated SNPs promote LUBAC formation and the ubiquitin E3 ligase activity. Although the exact mechanism by which these SNPs affect LUBAC activity will require structural studies, their position in the available RNF31–RBCK1 crystal structure offers some insight. The RNF31 Q622L mutant is located in an unusual, bent α-helical region that makes direct contact with RBCK1 (23). On the basis of the analysis of other proteins (35), the proline residue at position 619 would be predicted to create a 26° bend between the two adjacent α-helices, but the observed angle in RNF31 is approximately 53°. This acute bend seems to be reinforced by electrostatic interactions between glutamic acid 618 (E618) and both Q622 and arginine 621. The nonpolar leucine residue introduced by the Q622L polymorphism should not interact with E618 in this way, allowing the carboxy-terminal α-helix to adopt an alternative orientation with respect to RBCK1 and potentially promoting the observed stronger interaction between RNF31 and RBCK1. The Q584H polymorphism, which was functionally weaker than Q622L, is nonetheless located in the α6-helix that makes contact with the bent α8/α9-helix containing Q622. Q584 sits at a sharp kink between the α6- and α7-helices, making contact with E591. Conceivably, the introduction of a histidine at position 584 could alter this arrangement and indirectly affect the interaction of the α8/α9-helix with RBCK1.

The present study describes a new role for LUBAC in the CBM complex, whereby LUBAC promotes MALT1 cleavage of A20 during BCR signaling, presumably contributing to greater IKK activity. LUBAC coimmunoprecipitated with MALT1 in ABC DLBCL cells and was required for full IKK activity following BCR cross-linking in a GCB DLBCL line. The association of MALT1 with its substrate A20 was enhanced by the expression of RNF31, and mutant RNF31 isoforms that promote greater LUBAC activity increased MALT1/A20 association. One model to explain these observations would be that LUBAC-mediated ubiquitin of protein(s) in the CBM complex attracts A20, owing to the ability of A20 to bind to linear ubiquitin (28), thereby increasing the access of MALT1 to its substrate. Although neither MALT1 nor CARD11 was detectably modified by linear ubiquitin, IKK itself could attract A20 to the CBM complex, as it was heavily modified by linear ubiquitin and is an integral component of this complex in ABC DLBCL cells (6). Alternatively, LUBAC-mediated ubiquitination in the CBM complex could increase the intrinsic proteolytic activity of MALT1, by an unknown mechanism.

Previous studies have implicated LUBAC in various signaling pathways that engage NF-κB, including those triggered by TNF, IL-1β, and CD40 ligand (CD40L; refs. 11, 12, 14, 15). In the TNFR1-mediated pathway, LUBAC is recruited to the receptor complex in a TRAF2/c-IAP1/c-IAP2–dependent fashion by binding to c-IAP1/2–generated ubiquitin linkages (15). Thus, it is likely that LUBAC is recruited to the CBM complex via polyubiquitin chains attached to one or more subunits in this complex. Indeed, the CBM subunits MALT1 and BCL10 are modified by ubiquitin following T-cell receptor stimulation (36–38). Moreover, MALT1 is monoubiquitinated in ABC DLBCL cells, and this modification promotes their survival (38). Chronic active BCR signaling in ABC DLBCL presumably stimulates MALT ubiquitination in ABC DLBCL, but the mechanism by which this occurs remains to be elucidated.

Targeting the RNF31–RBCK1 interface using stapled α-helical RNF31 peptides specifically kills ABC DLBCL cells, supporting the development of LUBAC inhibitors for the therapy of ABC DLBCL. Despite the oncogenic role of many E3 ligases in cancer, small-molecule inhibitors of the active site of these enzymes have not yet emerged as therapies. Our work suggests that focusing small-molecule screens on the RNF31–RBCK1 interface might be a useful strategy. LUBAC inhibitors would be expected to have direct cytotoxic effects on the malignant ABC DLBCL cells but also, by inhibiting NF-κB, sensitize these cells to the apoptotic effects of conventional chemotherapeutic agents (32). Indeed, the RNF31 stapled peptides sensitized ABC DLBCL cells to etoposide. Our studies also suggest that LUBAC inhibitors should sensitize ABC DLBCLs to targeted agents that affect the NF-κB pathway, including ibrutinib and lenalidomide. When contemplating LUBAC as a therapeutic target, it is important to emphasize that LUBAC seems to participate in some, but not all, signaling events that activate IKK and NF-κB. In particular, mice with mutations that disrupt LUBAC components do not phenocopy mice with loss of IKK-β or -γ, which are characterized by massive liver apoptosis during development (11–14). Likewise, a rare inherited human immunodeficiency disease caused by loss-of-function RBCK1 mutations was characterized by loss of NF-κB activation in some cell types but gain of NF-κB activity in others (39). Thus, drugs targeting LUBAC would have a different spectrum of activities than IKK-β inhibitors. Given the presumed importance of BCR and TLR signaling in autoimmune/inflammatory diseases (40), LUBAC inhibitors might prove useful beyond ABC DLBCL.

Finally, our study demonstrates the value of interrogating rare germline polymorphisms for their role in cancer. Recent examples of rare germline polymorphisms contributing to tumorigenesis have been described in neuroblastoma and melanoma (41–44). These alleles are difficult to discover by standard genome-wide association methods, but can confer a high relative risk, as is the case for the RNF31 polymorphisms in ABC DLBCL. Our study highlights the importance of functional analysis in evaluating the contribution of rare SNPs to disease pathogenesis.

Patient Samples, PCR Amplification, and Sanger Sequencing

Tumor biopsy specimens before treatment were obtained from 302 patients with de novo DLBCL, which had previously been classified by gene expression profiling, and 116 patients with follicular lymphoma (FL), 75 patients with chronic lymphocytic leukemia, and 68 patients with Hodgkin lymphoma. All samples were studied according to a protocol approved by the National Cancer Institute Institutional Review Board. Genomic DNA from patient samples was extracted with the DNeasy Tissue Kit (Qiagen) according to the manufacturer's instructions. The primers used to amplify RNF31 exon 10 are RNF31_E10_F, 5′-CTGGGCTGGGTGCCTTTTCCTGTCAGG-3′ and RNF31_E10_R, 5′-GAGTAATTCTTGGACCAGGTATCG-3′. The PCR products were purified using the MinElute 96 UF PCR Purification Kit (Qiagen) and subsequently sequenced using the BigDye sequencing system (Applied Biosystems) from both strands.

Cell Culture

ABC- and GCB-derived DLBCL cell lines BJAB, HT, HBL1, DLBCL2, TMD8, OYB, SUDHL2, and TK were grown in RPMI-1640 medium (Invitrogen) + 10% FBS (Hyclone, Defined) + penicillin/streptomycin (Invitrogen); OCI-Ly3, OCI-Ly10, OCI-Ly8, and OCI-Ly19 cell lines were grown in Iscove's Modified Dulbecco's Medium (IMDM; Invitrogen) + 20% human serum + pen/strep (Invitrogen). All cell lines were grown to log phase at 37°C, 5% CO2 when experiments started. All cell lines had previously been modified to express an ecotropic retroviral receptor and a fusion protein of the Tet repressor and the blasticidin resistance gene, as described previously (5).

Cell Lines

ABC and GCB DLBCL cell lines were obtained from the following sources: Martin Dyer (University of Leicester, Leicester, United Kingdom; HBL1; ref. 45), Hans Messner (University of Toronto, Toronto, Canada; OCI-Ly3, OCILy8, OCI-Ly10, and OCI-Ly19; ref. 46), Shuji Tohda (Tokyo Medical and Dental University, Tokyo, Japan; TMD8; ref. 47), Momoko Nishikori (Kyoto University, Kyoto, Japan; OYB, DLBCL2; ref. 48), the Japanese Collection of Research Bioresources (JCRB) cell bank (TK; http://cellbank.nibio.go.jp), the American Type Culture Collection (HT, SUDHL2; http://www.atcc.org), and DMSZ (BJAB; http://www.dsmz.de). Cell lines have been characterized extensively by gene expression profiling (1) and cancer gene resequencing (6, 7, 49).

Retroviral Vectors and Transduction for shRNA Expression

The retroviral vectors for short hairpin RNA (shRNA) expression were described previously (8). In brief, the shRNA oligos were constructed into a pMSCV-based retroviral vector (pRSMX_Puro) with constitutive expression of a puromycin resistance marker fused with GFP. The inducible expression of shRNA was released after binding of the bacterial tetracycline repressor by doxycycline (50 ng/mL). For retroviral production, shRNA constructs were mixed with a mutant ecotropic envelope-expressing plasmid pHIT/EA6 × 3* and gag-pol–expressing plasmid pHIT60 and were transfected into 293T cells using the Lipofectamine 2000 reagent (Invitrogen) as described previously (5). The produced retroviral was used to infect doxycycline-inducible lymphoma cells in the presence of 8-μg/mL polybrene in a single spin infection, and puromycin was used to select for stable integrants over 6 days.

shRNA Sequences

Sequences for control- and target-specific shRNAs are as follows: control shRNA (CTCTCAACCCTTTAAATCTGA), MYC shRNA (CGATTCCTTCTAACAGAAATG), RNF31 shRNA #3 (GCCAGAGCTGTACCTTTGAGA), RNF31 shRNA #10 (GAAGACAAGGTTGAAGATGAT), SHARPIN shRNA #4 (GGACGCTTGTTTCCCCCATCA), and SHARPIN shRNA #6 (GAGCGCAGCCTTGCCTCTTAC).

shRNA Toxicity and Complementation Assays

The toxicity assay of shRNA was described previously (5). In brief, 2 days after infection with a retrovirus-expressing shRNA and GFP, the fraction of GFP-positive live cells was measured by flow cytometry. Doxycycline was then added to induce shRNA expression, and the fraction of GFP-positive live cells was measured at various time points during subsequent culture. The GFP-positive fraction from the test shRNA cultures was normalized to the GFP-positive fraction on day 0.

Retroviral construct used for ectopic expression of WT and mutant RNF31 was described previously (8). In brief, the retroviral vector for inducible cDNA expression was pMSCV based, with the cDNA expressed from a doxycycline-inducible cytomegalovirus (CMV) promoter in which a binding site for the bacterial tetracycline repressor is inserted at the transcription start site (derived from pCDNA4/TO; Invitrogen). The Myc-tagged human RNF31 cDNA was described previously (11) and was cloned into retroviral vector for ectopic expression. RNF31 Q584H and Q622L mutagenesis was performed with the QuikChange Kit from Stratagene and verified by Sanger sequencing. For the RNF31 shRNA rescue retroviral constructs, additional mutations were introduced to RNF31 shRNA-targeted sequence with the primers:

  • Forward, p-GCTGTTGGAAGACAAAGTAGAGGATGATATGCTGC;

  • Reverse, p-GCAGCATATCATCCTCTACTTTGTCTTCCAACAGC.

Antibody and Reagents

The antibody against linear ubiquitin chains and RBCK1 was described previously (13). Other antibodies were purchased as follows: anti–IKK-β, anti–phospho-IKK-β, anti-IκBα, anti–phospho-IκBα, and anti-CARD11 from Cell Signaling Technology; anti-ubiquitin (P4D1), polyclonal anti–NEMO/IKK-γ (FL-419), anti–β-actin, anti-IRAK1, anti-MALT1, and anti-A20 from Santa Cruz Biotechnology; anti-RNF31, anti-SHARPIN, and anti–Myc-tag from Abcam; monoclonal anti–NEMO/IKK-γ from BD Pharmingen; and anti-human IgM from Jackson ImmunoResearch Laboratories. Isotype control antibodies were obtained from the same company as each experimental antibody. Secondary horseradish peroxidase–conjugated antibodies were obtained from GE Healthcare.

The Myc-tag elution peptide and etoposide were obtained from Sigma. The IMiD compound lenalidomide was obtained from Celgene Corporation. The IKK-β inhibitor MLN120B was obtained from Millennium Pharmaceuticals. The BTK inhibitor ibrutinib was obtained from Pharmacyclics, Inc. The MALT1 inhibitor Z-VRPR-fmk was obtained from Enzo Life Sciences. Tissue culture–grade dimethyl sulfoxide (DMSO) vehicle control was obtained from Sigma-Aldrich.

Western Blotting

Cell pellets were lysed in modified radioimmunoprecipitation assay (RIPA) buffer (50 mmol/L Tris–HCl pH 7.5, 150 mmol/L NaCl, 1% NP40, 0.25% deoxycholic acid, and 1 mmol/L EDTA) supplemented with protease inhibitor tablet and phosphatase inhibitor tablet (Roche), 1 mmol/L dithiothreitol (DTT), 1 mmol/L Na3VaO4, and 1 mmol/L phenylmethylsulfonylfluoride (PMSF). Protein concentration was measured by the BCA Protein Assay Kit (Thermo Scientific). Total proteins were separated on 4% to 12% SDS-PAGE gels and transferred to nitrocellulose membranes.

NEMO Immunoprecipitation and Ubiquitination

For NEMO ubiquitination, cells were boiled for 15 minutes in 1% SDS before immunoprecipitation. Boiled lysates were diluted to 0.1% SDS with a modified RIPA buffer [50 mmol/L Tris–HCl pH 7.5, 150 mmol/L NaCl, 1% NP40, 0.25% deoxycholic acid, 1 mmol/L EDTA, supplemented with protease inhibitors, and 5 mmol/L N-ethylmaleimide (Sigma)]. Cleared lysates were incubated overnight with polyclonal anti–NEMO/IKK-γ antibody (Santa Cruz Biotechnology FL-419). Immunoprecipitates were washed five times with lysis buffer, separated by SDS-PAGE, transferred to nitrocellulose, and analyzed by immunoblotting with an anti-ubquitin antibody. The method for the detection of NEMO linear polyubiquitination was described previously (13). In brief, cells were boiled for 15 minutes in 1% SDS and then diluted to 0.1% SDS with a modified RIPA buffer. Cleared lysates were subjected to immunoprecipitation with an anti-NEMO monoclonal antibody (BD Pharmingen). Immunoprecipitates were separated on 4% to 12% SDS-PAGE gels and transferred to nitrocellulose membranes, autoclaved with distilled water at 121°C for 30 minutes, and then autoclaved again for 15 minutes without water. Membranes were analyzed by immunoblotting with an anti-linear ubiquitin antibody.

Coimmunoprecipitation

Cells were lysed in an endogenous lysis buffer (20 mmol/L Tris–HCl pH 7.6, 150 mmol/L NaCl, 1 mmol/L EDTA, 1% Triton X-100, 30 mmol/L NaF, and 2 mmol/L sodium pyrophosphate) supplemented with complete protease inhibitor cocktail (Roche), phosphatase inhibitor tablet (Roche), 1 mmol/L DTT, 1 mmol/L Na3VaO4, and 1 mmol/L PMSF. Cleared lysates were incubated overnight with polyclonal anti-MALT, anti-IRAK1, and control antibodies. Immunoprecipitates were washed five times with 0.5 mol/L NaCl lysis buffer, separated by SDS-PAGE, transferred to nitrocellulose, and analyzed by immunoblotting.

E3 Ubiquitin Ligase Assay and In Vitro Ubiquitination Assay

Engineered HBL1 lines induced to express various isoforms of Myc-tagged RNF31 were lysed in an endogenous lysis buffer (20 mmol/L Tris–HCl pH 7.6, 150 mmol/L NaCl, 1 mmol/L EDTA, 1% Triton X-100, 30 mmol/L NaF, and 2 mmol/L sodium pyrophosphate) supplemented with complete protease inhibitor cocktail (Roche), phosphatase inhibitor tablet (Roche), 1 mmol/L DTT, 1 mmol/L Na3VaO4, and 1 mmol/L PMSF. Cleared lysates were incubated overnight with an anti-Myc antibody. Immunoprecipitates were washed five times with lysis buffer, eluted with Myc-specific peptides, and subjected to E3 ubiquitin ligase assay using an E3LITE customizable ubiquitin ligase kit obtained from Life Sensors, following the manufacturer's instructions.

For the in vitro ubiquitination assay, the immunoprecipitated LUBAC complex was washed five times with lysis buffer, eluted with Myc-specific peptides, and resuspended in 40 μL of 20 mmol/L Tris–HCl pH 7.5, 2 mmol/L DTT, 0.1 μmol/L UBE1, 0.4 μmol/L UBCH5C, 10 μmol/L ubiquitin, 5 mmol/L MgCl2, and 2 mmol/L ATP. Reaction mixtures were incubated for 1 hour at 30°C and stopped by boiling for 10 minutes with SDS. The formation of linear-ubiquitin chains was analyzed by immunoblotting with an anti-ubiquitin antibody.

NF-κB p65 DNA-Binding ELISA

NF-κB p65 DNA-binding activity was measured using a TransAM NF-κB p65 ELISA kit obtained from Active Motif, following the manufacturer's instructions.

IkB Kinase Activity Reporter Assay

The assay for IκB kinase activity using the IκBα-Photinus luciferase reporter has been described previously (6). In brief, stable clones of TMD8 were constructed with vectors to express a fusion protein between IκBα and Photinus luciferase (from pGL3; Promega) as the reporter, and Renilla luciferase (from phRL-TK; Promega) for normalization. The ratio of IκBα-Photinus to Renilla luminescence was measured by the Dual-Glo Luciferase Assay System (Promega), and was normalized to that in untreated or uninduced controls.

NF-κB Reporter Assays

NF-κB transcriptional reporter ABC DLBCL lines were generated by transduction with lentiviral particles containing an inducible NF-κB responsive luciferase reporter construct (SA Biosciences) and selected with puromycin. Luciferase activity was measured using the Dual Luciferase Reporter Assay System (Promega) on a Microtiter Plate Luminometer (Dyn-Ex Technologies).

Flow Cytometry

Flow cytometry for NF-κB activation was performed 3 days after transgene infection. CD83 expression in BJAB cells was determined by staining with an anti-human CD83 antibody (BioLegend).

Peptide Synthesis

Peptide synthesis, olefin metathesis, FITC derivatization, reverse-phase high-performance liquid chromatography (HPLC) purification, and amino acid analysis were performed as described previously (50).

Circular Dichroism Spectroscopy

Peptides (dry, powder form) were dissolved in H2O to prepare 50 μmol/L solutions. The spectra were obtained on a Jasco J-715 spectropolarimeter at 20°C. The spectra were collected using a 0.1-cm pathlength quartz cuvette with the following measurement parameters: wavelength, 185–255 nm; step resolution, 0.2 nm; speed, 20 nm/min; accumulations, 3; and bandwidth, 1 nm.

Fluorescence Polarization Assay and Competition Assay

For binding assays, FITC-peptide (LT = 14.1 nmol/L) was incubated with a broad range of GST-RBCK1 concentrations in 50 mmol/L Tris, 150 mmol/L NaCl, pH 8.0 at 4°C. Binding activity was measured by fluorescence polarization on a SpectraMax M5 Microplate Reader (Molecular Devices) in a black, polystyrene, nontreated, 96-well plate (Costar, Corning Inc.) at 20 minutes. Kd values were determined by nonlinear regression analysis of dose–response curves using Prism GraphPad software v 6.0. Each data point represents the average of an experimental condition performed in at least triplicate. For competition assays, FITC-RNF31-N Q2L peptide (14.1 nmol/L) was combined with a serial dilution of unlabeled, Ac-RNF31 N-WT, or Ac-RNF31 N-Q622L peptide, followed by the addition of GST-RBCK1 protein (700 nmol/L). IC50 values for FITC-peptide displacement were calculated by nonlinear regression analysis using Prism software (GraphPad).

RBCK1 Recombinant Protein

The codon-optimized cDNA for Escherichia coli expression of full-length human RBCK1 was a kind gift of Dr. Titia K. Sixma (Division of Biochemistry, The Netherlands Cancer Institute, Amsterdam, the Netherlands). For expression, the transformed E. coli Bl21 (DE3) pLysS cells were inducted with 0.8 mmol/L isopropyl-l-thio-B-[scap)d[r]-galactopyranoside (IPTG) and 0.2 mmol/L ZnSO4 for 8 hours at 22°C. Cells were lysed by using the B-PER bacterial protein extraction reagent (Thermo Scientific), and full-length human RBCK1 was purified with a Pierce GST spin purification kit (Thermo Scientific).

Confocal Microscopy

Images were acquired using a Zeiss LSM510 Meta laser scanning confocal microscope equipped with a 40× C-Aprochromat (N.A., 1.2) objective lens, transmitted light detector, and differential interference contrast optical components. Confocal fluorescence images were collected with consistent detector settings for all samples, including 0.11-μm X–Y pixel size, 1.5-μm optical slice thickness, and 4× frame averaging. The final images were exported as TIFF files and arranged into figures using Adobe Photoshop (v.9.0). The brightness and contrast was adjusted equally for all images, using histogram stretching and adjustment of gamma to a value of 1.10.

Cell Viability (MTS) Assay

Cells were plated in triplicate at a density of 10,000 cells per well in 96-well plates. Cell viability after indicated treatments was assayed by adding MTS reagents (Promega), incubating for 1 hour, and measuring the amount of 490-nm absorbance using a 96-well plate reader. The background was subtracted using a media-only control.

Statistical Analysis

All experiments presented have been repeated and results reproduced. Where possible, error bars or P values are shown to indicate statistical significance. For the SNP enrichment analysis, two-sided P values for differences in prevalence were calculated using a Fisher exact test. Confidence levels for the enrichment values were calculated using a normal approximation to the log enrichment. CIs for prevalence estimates were calculated using the Clopper–Pearson method.

W.C. Chan is a consultant/advisory board member of the Lymphoma Research Foundation. No potential conflicts of interest were disclosed by the other authors.

Conception and design: Y. Yang, F. Bernal, L.M. Staudt

Development of methodology: Y. Yang, J. Mitala, W. Xiao, K. Iwai, F. Bernal, L.M. Staudt

Acquisition of data (provided animals, acquired and managed patients, provided facilities, etc.): Y. Yang, R. Schmitz, J. Mitala, A. Rosenwald, G. Ott, R.D. Gascoyne, J.M. Connors, L.M. Rimsza, E. Campo, E.S. Jaffe, J. Delabie, E.B. Smeland, R.R. Tubbs, J.R. Cook, W.C. Chan, A. Wiestner, M.J. Kruhlak, F. Bernal

Analysis and interpretation of data (e.g., statistical analysis, biostatistics, computational analysis): Y. Yang, R. Schmitz, J. Mitala, A. Whiting, W. Xiao, G.W. Wright, A. Rosenwald, F. Bernal, L.M. Staudt

Writing, review, and/or revision of the manuscript: Y. Yang, R. Schmitz, A. Rosenwald, G. Ott, R.D. Gascoyne, J.M. Connors, L.M. Rimsza, E. Campo, J. Delabie, E.B. Smeland, R.M. Braziel, J.R. Cook, D.D. Weisenburger, W.C. Chan, K. Iwai, F. Bernal, L.M. Staudt

Administrative, technical, or material support (i.e., reporting or organizing data, constructing databases): A. Whiting, M. Ceribelli, H. Zhao, Y. Yang, W. Xu, A. Rosenwald, W.C. Chan, F. Bernal

Study supervision: F. Bernal, L.M. Staudt

Pathology review: R.M. Braziel

The authors thank the patients for their participation. This study was conducted under the auspices of the Lymphoma/Leukemia Molecular Profiling Project (LLMPP).

This research was supported by the Intramural Research Program of the NIH, National Cancer Institute, Center for Cancer Research, and by an NCI SPECS grant (UO1-CA 114778) to L.M. Staudt. R. Schmitz was supported by the Dr. Mildred Scheel Stiftung für Krebsforschung (Deutsche Krebshilfe).

1.
Alizadeh
AA
,
Eisen
MB
,
Davis
RE
,
Ma
C
,
Lossos
IS
,
Rosenwald
A
, et al
Distinct types of diffuse large B-cell lymphoma identified by gene expression profiling
.
Nature
2000
;
403
:
503
11
.
2.
Shaffer
AL
 III
,
Young
RM
,
Staudt
LM
. 
Pathogenesis of human B cell lymphomas
.
Annu Rev Immunol
2012
;
30
:
565
610
.
3.
Davis
RE
,
Brown
KD
,
Siebenlist
U
,
Staudt
LM
. 
Constitutive nuclear factor kappa B activity is required for survival of activated B Cell-like diffuse large B cell lymphoma cells
.
J Exp Med
2001
;
194
:
1861
74
.
4.
Yang
Y
,
Shaffer
AL
 III
,
Emre
NC
,
Ceribelli
M
,
Zhang
M
,
Wright
G
, et al
Exploiting synthetic lethality for the therapy of ABC diffuse large B cell lymphoma
.
Cancer Cell
2012
;
21
:
723
37
.
5.
Ngo
VN
,
Davis
RE
,
Lamy
L
,
Yu
X
,
Zhao
H
,
Lenz
G
, et al
A loss-of-function RNA interference screen for molecular targets in cancer
.
Nature
2006
;
441
:
106
10
.
6.
Lenz
G
,
Davis
RE
,
Ngo
VN
,
Lam
L
,
George
TC
,
Wright
GW
, et al
Oncogenic CARD11 mutations in human diffuse large B cell lymphoma
.
Science
2008
;
319
:
1676
9
.
7.
Davis
RE
,
Ngo
VN
,
Lenz
G
,
Tolar
P
,
Young
RM
,
Romesser
PB
, et al
Chronic active B-cell-receptor signalling in diffuse large B-cell lymphoma
.
Nature
2010
;
463
:
88
92
.
8.
Ngo
VN
,
Young
RM
,
Schmitz
R
,
Jhavar
S
,
Xiao
W
,
Lim
KH
, et al
Oncogenically active MYD88 mutations in human lymphoma
.
Nature
2011
;
470
:
115
9
.
9.
Lam
LT
,
Wright
G
,
Davis
RE
,
Lenz
G
,
Farinha
P
,
Dang
L
, et al
Cooperative signaling through the signal transducer and activator of transcription 3 and nuclear factor-{kappa}B pathways in subtypes of diffuse large B-cell lymphoma
.
Blood
2008
;
111
:
3701
13
.
10.
Chen
ZJ
. 
Ubiquitination in signaling to and activation of IKK
.
Immunol Rev
2012
;
246
:
95
106
.
11.
Tokunaga
F
,
Sakata
S
,
Saeki
Y
,
Satomi
Y
,
Kirisako
T
,
Kamei
K
, et al
Involvement of linear polyubiquitylation of NEMO in NF-kappaB activation
.
Nat Cell Biol
2009
;
11
:
123
32
.
12.
Ikeda
F
,
Deribe
YL
,
Skanland
SS
,
Stieglitz
B
,
Grabbe
C
,
Franz-Wachtel
M
, et al
SHARPIN forms a linear ubiquitin ligase complex regulating NF-kappaB activity and apoptosis
.
Nature
2011
;
471
:
637
41
.
13.
Tokunaga
F
,
Nakagawa
T
,
Nakahara
M
,
Saeki
Y
,
Taniguchi
M
,
Sakata
S
, et al
SHARPIN is a component of the NF-kappaB–activating linear ubiquitin chain assembly complex
.
Nature
2011
;
471
:
633
6
.
14.
Gerlach
B
,
Cordier
SM
,
Schmukle
AC
,
Emmerich
CH
,
Rieser
E
,
Haas
TL
, et al
Linear ubiquitination prevents inflammation and regulates immune signalling
.
Nature
2011
;
471
:
591
6
.
15.
Haas
TL
,
Emmerich
CH
,
Gerlach
B
,
Schmukle
AC
,
Cordier
SM
,
Rieser
E
, et al
Recruitment of the linear ubiquitin chain assembly complex stabilizes the TNF-R1 signaling complex and is required for TNF-mediated gene induction
.
Mol Cell
2009
;
36
:
831
44
.
16.
Hostager
BS
,
Fox
DK
,
Whitten
D
,
Wilkerson
CG
,
Eipper
BA
,
Francone
VP
, et al
HOIL-1L interacting protein (HOIP) as an NF-kappaB regulating component of the CD40 signaling complex
.
PLoS ONE
2010
;
5
:
e11380
.
17.
Hostager
BS
,
Kashiwada
M
,
Colgan
JD
,
Rothman
PB
. 
HOIL-1L interacting protein (HOIP) is essential for CD40 signaling
.
PLoS ONE
2011
;
6
:
e23061
.
18.
Zak
DE
,
Schmitz
F
,
Gold
ES
,
Diercks
AH
,
Peschon
JJ
,
Valvo
JS
, et al
Systems analysis identifies an essential role for SHANK-associated RH domain-interacting protein (SHARPIN) in macrophage Toll-like receptor 2 (TLR2) responses
.
Proc Natl Acad Sci U S A
2011
;
108
:
11536
41
.
19.
Sasaki
Y
,
Sano
S
,
Nakahara
M
,
Murata
S
,
Kometani
K
,
Aiba
Y
, et al
Defective immune responses in mice lacking LUBAC-mediated linear ubiquitination in B cells
.
EMBO J
2013
;
32
:
2463
76
.
20.
Lenz
G
,
Wright
G
,
Dave
SS
,
Xiao
W
,
Powell
J
,
Zhao
H
, et al
Stromal gene signatures in large-B-cell lymphomas
.
N Engl J Med
2008
;
359
:
2313
23
.
21.
1000 Genomes Project Consortium
Abecasis
GR
,
Auton
A
,
Brooks
LD
,
DePristo
MA
,
Durbin
RM
, et al
An integrated map of genetic variation from 1,092 human genomes
.
Nature
2012
;
491
:
56
65
.
22.
Tennessen
JA
,
Bigham
AW
,
O'Connor
TD
,
Fu
W
,
Kenny
EE
,
Gravel
S
, et al
Evolution and functional impact of rare coding variation from deep sequencing of human exomes
.
Science
2012
;
337
:
64
9
.
23.
Yagi
H
,
Ishimoto
K
,
Hiromoto
T
,
Fujita
H
,
Mizushima
T
,
Uekusa
Y
, et al
A non-canonical UBA–UBL interaction forms the linear-ubiquitin-chain assembly complex
.
EMBO Rep
2012
;
13
:
462
8
.
24.
Lam
LT
,
Davis
RE
,
Pierce
J
,
Hepperle
M
,
Xu
Y
,
Hottelet
M
, et al
Small molecule inhibitors of IkappaB kinase are selectively toxic for subgroups of diffuse large B-cell lymphoma defined by gene expression profiling
.
Clin Cancer Res
2005
;
11
:
28
40
.
25.
Hailfinger
S
,
Lenz
G
,
Ngo
V
,
Posvitz-Fejfar
A
,
Rebeaud
F
,
Guzzardi
M
, et al
Essential role of MALT1 protease activity in activated B cell-like diffuse large B-cell lymphoma
.
Proc Natl Acad Sci U S A
2009
;
106
:
19946
51
.
26.
Ferch
U
,
Kloo
B
,
Gewies
A
,
Pfander
V
,
Duwel
M
,
Peschel
C
, et al
Inhibition of MALT1 protease activity is selectively toxic for activated B cell-like diffuse large B cell lymphoma cells
.
J Exp Med
2009
;
206
:
2313
20
.
27.
Tokunaga
F
,
Nishimasu
H
,
Ishitani
R
,
Goto
E
,
Noguchi
T
,
Mio
K
, et al
Specific recognition of linear polyubiquitin by A20 zinc finger 7 is involved in NF-kappaB regulation
.
EMBO J
2012
;
31
:
3856
70
.
28.
Verhelst
K
,
Carpentier
I
,
Kreike
M
,
Meloni
L
,
Verstrepen
L
,
Kensche
T
, et al
A20 inhibits LUBAC-mediated NF-kappaB activation by binding linear polyubiquitin chains via its zinc finger 7
.
EMBO J
2012
;
31
:
3845
55
.
29.
Smit
JJ
,
Monteferrario
D
,
Noordermeer
SM
,
van Dijk
WJ
,
van der Reijden
BA
,
Sixma
TK
. 
The E3 ligase HOIP specifies linear ubiquitin chain assembly through its RING-IBR-RING domain and the unique LDD extension
.
EMBO J
2012
;
31
:
3833
44
.
30.
Stieglitz
B
,
Morris-Davies
AC
,
Koliopoulos
MG
,
Christodoulou
E
,
Rittinger
K
. 
LUBAC synthesizes linear ubiquitin chains via a thioester intermediate
.
EMBO Rep
2012
;
13
:
840
6
.
31.
Walensky
LD
,
Kung
AL
,
Escher
I
,
Malia
TJ
,
Barbuto
S
,
Wright
RD
, et al
Activation of apoptosis in vivo by a hydrocarbon-stapled BH3 helix
.
Science
2004
;
305
:
1466
70
.
32.
Baldwin
AS
. 
Control of oncogenesis and cancer therapy resistance by the transcription factor NF-kappaB
.
J Clin Invest
2001
;
107
:
241
6
.
33.
McCool
KW
,
Miyamoto
S
. 
DNA damage-dependent NF-kappaB activation: NEMO turns nuclear signaling inside out
.
Immunol Rev
2012
;
246
:
311
26
.
34.
Niu
J
,
Shi
Y
,
Iwai
K
,
Wu
ZH
. 
LUBAC regulates NF-kappaB activation upon genotoxic stress by promoting linear ubiquitination of NEMO
.
EMBO J
2011
;
30
:
3741
53
.
35.
Barlow
DJ
,
Thornton
JM
. 
Helix geometry in proteins
.
J Mol Biol
1988
;
201
:
601
19
.
36.
Oeckinghaus
A
,
Wegener
E
,
Welteke
V
,
Ferch
U
,
Arslan
SC
,
Ruland
J
, et al
Malt1 ubiquitination triggers NF-kappaB signaling upon T-cell activation
.
EMBO J
2007
;
26
:
4634
45
.
37.
Sun
L
,
Deng
L
,
Ea
CK
,
Xia
ZP
,
Chen
ZJ
. 
The TRAF6 ubiquitin ligase and TAK1 kinase mediate IKK activation by BCL10 and MALT1 in T lymphocytes
.
Mol Cell
2004
;
14
:
289
301
.
38.
Pelzer
C
,
Cabalzar
K
,
Wolf
A
,
Gonzalez
M
,
Lenz
G
,
Thome
M
. 
The protease activity of the paracaspase MALT1 is controlled by monoubiquitination
.
Nat Immunol
2013
;
14
:
337
45
.
39.
Boisson
B
,
Laplantine
E
,
Prando
C
,
Giliani
S
,
Israelsson
E
,
Xu
Z
, et al
Immunodeficiency, autoinflammation and amylopectinosis in humans with inherited HOIL-1 and LUBAC deficiency
.
Nat Immunol
2012
;
13
:
1178
86
.
40.
Green
NM
,
Marshak-Rothstein
A
. 
Toll-like receptor driven B cell activation in the induction of systemic autoimmunity
.
Semin Immunol
2011
;
23
:
106
12
.
41.
Janoueix-Lerosey
I
,
Lequin
D
,
Brugieres
L
,
Ribeiro
A
,
de Pontual
L
,
Combaret
V
, et al
Somatic and germline activating mutations of the ALK kinase receptor in neuroblastoma
.
Nature
2008
;
455
:
967
70
.
42.
Mosse
YP
,
Laudenslager
M
,
Longo
L
,
Cole
KA
,
Wood
A
,
Attiyeh
EF
, et al
Identification of ALK as a major familial neuroblastoma predisposition gene
.
Nature
2008
;
455
:
930
5
.
43.
Yokoyama
S
,
Woods
SL
,
Boyle
GM
,
Aoude
LG
,
MacGregor
S
,
Zismann
V
, et al
A novel recurrent mutation in MITF predisposes to familial and sporadic melanoma
.
Nature
2011
;
480
:
99
103
.
44.
Bertolotto
C
,
Lesueur
F
,
Giuliano
S
,
Strub
T
,
de Lichy
M
,
Bille
K
, et al
A SUMOylation-defective MITF germline mutation predisposes to melanoma and renal carcinoma
.
Nature
2011
;
480
:
94
8
.
45.
Abe
M
,
Nozawa
Y
,
Wakasa
H
,
Ohno
H
,
Fukuhara
S
. 
Characterization and comparison of two newly established Epstein-Barr virus-negative lymphoma B-cell lines. Surface markers, growth characteristics, cytogenetics, and transplantability
.
Cancer
1988
;
61
:
483
90
.
46.
Tweeddale
ME
,
Lim
B
,
Jamal
N
,
Robinson
J
,
Zalcberg
J
,
Lockwood
G
, et al
The presence of clonogenic cells in high-grade malignant lymphoma: a prognostic factor
.
Blood
1987
;
69
:
1307
14
.
47.
Tohda
S
,
Sato
T
,
Kogoshi
H
,
Fu
L
,
Sakano
S
,
Nara
N
. 
Establishment of a novel B-cell lymphoma cell line with suppressed growth by gamma-secretase inhibitors
.
Leuk Res
2006
;
30
:
1385
90
.
48.
Maesako
Y
,
Uchiyama
T
,
Ohno
H
. 
Comparison of gene expression profiles of lymphoma cell lines from transformed follicular lymphoma, Burkitt's lymphoma and de novo diffuse large B-cell lymphoma
.
Cancer Sci
2003
;
94
:
774
81
.
49.
Schmitz
R
,
Young
RM
,
Ceribelli
M
,
Jhavar
S
,
Xiao
W
,
Zhang
M
, et al
Burkitt lymphoma pathogenesis and therapeutic targets from structural and functional genomics
.
Nature
2012
;
490
:
116
20
.
50.
Bernal
F
,
Tyler
AF
,
Korsmeyer
SJ
,
Walensky
LD
,
Verdine
GL
. 
Reactivation of the p53 tumor suppressor pathway by a stapled p53 peptide.
J Am Chem Soc
2007
;
129
:
2456
7
.