The rapid increase in the incidence of the B cell non-Hodgkin's lymphomas (NHL) and improved understanding of the mechanisms involved in their development renders timely a review of the theoretical and practical aspects of molecular abnormalities in B cell NHL.

In Section I, Dr. Macintyre addresses the practical aspects of the use of molecular techniques for the diagnosis and therapeutic management of patients with B cell NHL. While detection of clonal Ig rearrangements is widely used to distinguish reactive from malignant lymphoproliferative disorders, molecular informativity is variable. The relative roles of cytogenetic, molecular and immunological techniques in the detection of genetic abnormalities and their protein products varies with the clinical situation. Consequently, the role of molecular analysis relative to morphological classification is evolving. Integrated diagnostic services are best equipped to cope with these changes. Recent evidence that large scale gene expression profiling allows improved prognostic stratification of diffuse large cell lymphoma suggests that the choice of diagnostic techniques will continue to change significantly and rapidly.

In Section II, Dr. Willerford reviews current understanding of the mechanisms involved in immunoglobulin (Ig) gene rearrangement during B lymphoid development and the way in which these processes may contribute to Ig-locus chromosome translocations in lymphoma. Recent insights into the regulation of Ig gene diversification indicate that genetic plasticity in B lymphocytes is much greater than previously suspected. Physiological genomic instability, which may include isotype switching, recombination revision and somatic mutation, occurs in germinal centers in the context of immune responses and may explain longstanding clinical observations that link immunity and lymphoid neoplasia. Data from murine models and human disorders predisposing to NHL have been used to illustrate these issues.

In Section III, Dr. Morris reviews the characteristics and consequences of deregulation of novel “proto-oncogenes” involved in B cell NHL, including PAX5 (chromosome 9p 13), BCL8 (15q11-q13), BCL9, MUC1, FcγRIIB and other 1q21-q22 genes and BCL10 (1p22). The AP12-MLT/MALT1 [t(11;18)(q21;q21)] fusion transcript is also described.

Elizabeth Macintyre, M.D.*

*

Laboratoire d'Hematologie, Tour Pasteur, Hôpital Necker Enfants Malades, Rue de Sevres, Paris 75743, France

Molecular techniques are of increasing practical importance in the analysis of non-Hodgkin's lymphomas (NHL), both for diagnostic purposes and in order to accurately evaluate prognosis. In addition, clonal or disease-specific abnormalities can provide markers for the detection of minimal residual disease. An increasing number of largely complementary techniques, including Southern blotting, PCR amplification of DNA or RNA and fluorescent in-situ hybridization (FISH), are responsible for these advances. In view of the number of investigational techniques available it has become necessary to evaluate their relative roles and to assess their contribution with respect to classical techniques, including morphology, immunology and karyotype analysis. This presentation will be limited to evaluation of cellular events that directly or indirectly reflect the oncogenic process; molecular detection of viral agents that play a role in lymphomagenesis will not be discussed. The former can be divided into oncogenic events, as detailed in Section III, and physiological Ig and TCR rearrangements that provide useful clonal markers but are not directly involved in lymphoma development, other than as described in Section II.

Clonal Markers

The vast majority of NHL have undergone physiological, clonal immunoglobulin (Ig) or T Cell Receptor (TCR) rearrangements (Section II). The identification of a clonal Ig/TCR rearrangement (lymphoid clonality) is therefore widely used at diagnosis and increasingly for follow-up. It probably represents the most frequently performed diagnostic molecular analysis in NHL.

Technical Aspects

Until recently, detection of lymphoid clonality was largely based on Southern blot analysis of predominantly IgH and TCRβ genes. The technique is based on detection of aberrantly sized DNA fragments following restriction endonuclease digestion and hybidization with labeled probes. Both partial D-J and complete V(D)J rearrangements are detected and somatic mutation does not prevent hybridization. Chromosomal translocations involving Ig/TCR are also detected. The limited repertoire of TCRγ and TCRδ genes means that it is possible to determine the V and J segments involved in a given clonal rearrangement based on the size of DNA fragments obtained but that it is difficult to distinguish minor clonal populations from polyclonal TCRγδ populations using a single V-J combination, since Southern analysis will not distinguish junctional diversity. The Southern technique is, however, long, technically demanding, requires a considerable amount of material (approximately 20μg) and, although its optimal sensitivity is approximately 5%, it is often less sensitive. It is therefore increasingly replaced by PCR analysis. Our current practice for detection of lymphoid clonality is to use Southern blotting to further investigate discordant or suspected false-negative IgH results and to evaluate potential IgH oligoclonality, which is poorly identified by consensus PCR strategies.

PCR detection of Ig/TCR V(D)J gene rearrangements exploits the fact that the CDR3 or N regions are clone specific and highly variable, both with regard to length and content. By using V- and J-specific or consensus primers and separation of PCR products on the basis of size it is possible to distinguish clonal from reactive populations. Amplification of a clonal lymphoid population generates a discrete band on electrophoresis whereas polyclonal rearrangements generate a smear of variably sized products. Amplification is performed from DNA, although RT-PCR amplification of Ig/TCR transcripts using V and C region primers, often with subsequent J segment fluorescent “run-off” analysis, is possible and has been widely exploited for functional repertoire analysis of mature lymphoid populations (Immunoscope).1  DNA is preferred for clonality studies since it will also detect rearrangements in immature populations that may not be transcribed and is more representative of the genomic rather than the functional repertoire. RNA analysis, however, renders accessible complex loci such as TCRα and TCRβ. Detection of partial IgH DH-JH rearrangements requires the use of DH consensus primers.2 

A variety of detection methods for Ig/TCR PCR products exist, with, in general, the sensitivity being proportional to the complexity. Detection is classically performed on non-denaturing polyacrylamide gel electrophoresis (PAGE) with ethidium bromide staining, often with prior heating of PCR products to encourage heteroduplex formation, with a sensitivity of approximately 10-1 to 10-2 (1-10%). This is increasingly being replaced by analysis of fluorescent PCR products on DNA fragment analysers (automated sequencers), allowing increased resolution and an approximately 1 log increase in sensitivity (10-2 to 10-3; 0.1-1%).3  It is also more user friendly than its radioactive predecessor (DNA fingerprinting).4  Family-specific amplification is more sensitive than consensus strategies since there is less competition for reagents from normal, polyclonal lymphocytes, but is less well adapted to initial diagnostic screening, unless incorporated in a multiplex format.5  Alternative strategies that allow separation of PCR products on the basis of nucleotide content as well as size, such as denaturing-gradient gel electrophoresis (DGGE),6  temperature-gradient gel electrophoresis (TGGE)7  or single-stranded conformational polymorphism (SSCP),8  have been described, particularly for analysis of TCRγ. The most sensitive and specific detection techniques involve sequencing of the clonal V-J junction and synthesis of an anti-junctional (AJO) or allele-specific oligonucleotide (ASO), which is then used a hybridization probe or clone specific primer.9,10,11  The sensitivity of such strategies is approximately 10-5, although this varies with the junctional sequence. While these strategies represent optimal specificity and sensitivity, they are labor intensive and do not permit detection of clonal evolution. It is therefore probable that they will be replaced by simpler strategies that provide an intermediate level of detection and that may be more easily applicable to large scale clinical follow-up. The arrival of real-time PCR quantification has already had a major impact on the detection and follow-up of oncogenic fusion transcripts. Adapting this type of technology to the detection of clonal V(D)J rearrangements is less simple, but possible.12,13 

IgH represents the most useful gene target for detecting B cell clonality since it rearranges early during B lymphoid development and demonstrates extensive junctional diversity. Because of the marked combinatorial diversity it is necessary to use consensus or family-specific primers. Consensus VH (FR1, FR2 and FR3) and JH (FR4) primers have been described, and family-specific primers are used either individually or in a multiplex format. The main limitation of IgH PCR is false-negative results, the incidence of which varies with the IgH PCR strategy and the pathological subtype as a function of the degree of receptor editing (somatic mutation, intraclonal variability, etc.; see Section II and 14 ).

TCRγ represents the most useful marker for T cell clonality since it is rearranged at an early stage of both TCR γδ and TCR αβ lineage cells and is easily accessible by PCR. Due to the relatively limited combinatorial diversity it is possible to detect all functional Vγ-Jγ rearrangements with a small panel of Vγ- and Jγ-specific primers, commonly in 1 to 3 multiplex reactions.5  False negative results due to failure to recognise the Vγ and Jγ segments involved are very rare if all Vγ and Jγ combinations are included. Junctional diversity is, however, limited, with an approximate CDR3 length of 20 nucleotides compared to 60 nucleotides for IgH. A further difference in the gaussian distribution of polyclonal TCRγ versus IgH PCR products results from the fact that TCRγ rearrangements only undergo selection of in frame rearrangements in TCR γδ-expressing cells, which represent a minority of circulating lymphocytes, and the PCR peaks generated therefore differ by only 1 bp (compared to 3 bp for IgH). A further problem in TCRγ PCR is the presence of normal conserved rearrangements with little junctional diversity (“canonical” TCRγ rearrangements that do not demonstrate N nucleotide additions) in DNA from circulating TCR γδ-expressing cells, which can mistakenly be interpreted as clonal bands in polyclonal populations.15  For all these reasons, the main limitation of detection of TCRγ rearrangements by PCR is the risk of false-positive results. This appears to be particularly problematic in the analysis of peripheral blood samples from older individuals. It is therefore extremely important to analyze TCRγ PCR products using either high resolution electrophoretic techniques or electrophoretic systems that allow separation on the basis of nucleotide content as well as size, and to be aware of the size and nature of canonical rearrangements, particularly those involving Vγ9-JγP (also known as Vγ2-Jγ1.2).15  It is not possible to distinguish polyclonal and clonal TCRγ rearrangements on classical agarose gel electrophoresis.

Clinical Applications at Diagnosis

Analysis of lymphoid clonality is frequently useful in the distinction of malignant vs. reactive lymphadenopathy in cases for which morphological and immunological analysis is inconclusive. Detection of apparently clonal proliferations in clinically benign conditions such as large granular lymphocytosis, lymphomatoid papulosis or Sjögren's syndrome illustrate, however, that clonality is not synonymous with malignancy,16, 17  and it is increasingly clear that dys/autoimmune disorders represent one end of the spectrum of lymphoproliferative disorders. Interpretation of molecular results, particularly for TCR analyses (see above), should therefore be undertaken in close collaboration with histopathologists and should take into account clinical, immunological and (cyto)genetic features.

“Illegitimate” rearrangements, by which TCR rearrangements occur in B lineage cases and vice versa, are frequent in ALL, probably reflecting the fact that the blasts are arrested at a “recombinase competent” stage of maturation arrest. In contrast, such rearrangements are rare in NHL. This may be due to the fact that oncogenic conversion occurs in a recombinase-incompetent cell or, more likely, that the chromosomal configuration of illegitimate targets are not in an accessible configuration. For these reasons, analysis of Ig and TCR genes can aid determination of T- or B-cell lineage, although this distinction is predominantly based on immunophenotyping. Co-existent clonal Ig and TCR rearrangements are, however, recognized in disorders such as angioimmunoblastic lymphadenopathy with dysplasia (AILD).18, 19  It is probable that they originate from different cell populations, as a consequence of the immune deregulation.

The incidence of informativity of IgH PCR in B-cell tumors varies with the stage of maturation, mainly as a function of their pre- or post-germinal center phenotype.14,20  It is lowest for follicular lymphoma (FL) and HIV-associated NHL and highest in CD5-expressing proliferations such as mantle cell lymphoma (MCL).21,22,23  For the reasons stated above, molecular informativity is rarely a problem with TCRγ PCR analysis. Since histological distinction of clonal and reactive T lymphocytes is often problematic, particularly in extranodal tissues such as skin or gut, the latter represent an increasingly common source of material transmitted to laboratories performing lymphoid clonality analyses.24,25 

Molecular Genetic Abnormalities

Technical aspects

Only general aspects pertaining to diagnostic use of molecular genetic markers will be discussed here. Experience with their practical use is largely restricted to markers that are sufficiently common, indicative of a particular diagnostic or therapeutic subgroup or useful for disease follow-up.

In molecular terms, chromosomal abnormalities or their sub-microscopic equivalents are of two general types: those in which the breakpoints occur within the involved genes, leading to the production of a fusion RNA transcript and a chimeric protein (Section III), and those that represent Ig/TCR rearrangement errors (Section II). The former lead to a qualitative change, insofar as the fusion protein is more or less tumor specific. They are relatively rare in B lineage NHL, compared to their acute leukemic counterparts, and Ig errors predominate. It is not clear whether these differences are due to a bias of identification or a real difference in oncogenic mechanisms. Qualitative abnormalities are usually detected by reverse transcriptase (RT) PCR from RNA, although DNA-based detection is possible for some cases. The best characterized example in NHL is the NPM-ALK fusion transcript,26,27  but the recently described AP12-MLT/MALT1 also falls into this category (see Section III).

Ig gene-associated translocations in B-NHL predominantly involve the IgH chain locus at 14q32 and, more rarely, Igκ and Igλ light chain loci at chromosomes 2p12 and 22q11, respectively. Errors of the recombination process lead to juxtapositioning of a proto-oncogene, with its consequent transcriptional deregulation and a quantitative change in its expression. The breakpoints occur outside the coding sequences, often at a distance of several kbp from the deregulated proto-oncogene. Genomic clustering facilitates detection of the translocation breakpoints by PCR from DNA, at least for those involving D-J breakpoints. Since the proto-oncogene-(D)-J junctional sequence demonstrates features identical to those seen in normal V(D)J rearrangement, these junctional sequences generate variably sized PCR products that can be used to distinguish similar translocations in different individuals and to identify contaminants.

Comparison of cytogenetic, molecular genetic and immunological analysis

Only karyotype analysis by classical morphological analysis of metaphase spreads or by multifluorescent SKY or M-FISH techniques28,29  provides an overall evaluation of the whole genome and directs further analysis. It is clear that the latter allow a substantial increase in resolution and capacity to identify abnormal chromosomes, with a consequent increase in the detection of abnormal karyotypes. In contrast, molecular techniques allow detection of specific abnormalities in situations where karyotype analysis is difficult, such as evaluation of archival tissue, when insufficient metaphases have been obtained, or for the detection of intrachromosomal abnormalities. Furthermore, it is increasingly recognized that molecular-positive, karyotype-negative cases are not uncommon, particularly if molecular cytogenetic techniques are not employed. They include cases with complex abnormalities, those with mitoses that are not representative of the tumor clone and those with sub-microscopic or “perfectly equilibrated” abnormalities. Notable examples of the latter category include TEL-AML1 [t(12;21)(p13;q22) translocation] in B cell lineage ALL30,31  and FGFR3-IgH [t(4;14)(p16;q32) translocation] in myeloma.32,33 

Use of whole chromosome FISH paints is appropriate for detection of numerical abnormalities, although these rarely have a practical impact in lymphoma management. Locus or rearrangement-specific probes can be used to aid diagnosis, especially for the detection of heterogeneous abnormalities such as MYC-IgH or BCL6 (3q27)34,35  rearrangements, which are difficult to exclude by Southern and even more difficult by PCR. The telomeric location of the IgH locus renders particularly difficult the detection of the partner chromosome, thus explaining the common 14q+ category of abnormalities. FISH analysis with IgH probes can help to characterize these cases.36  Specificity of detection can be improved by the use of dual color FISH.37  Probes that allow identification of abnormalities in interphase nuclei are preferable for screening purposes in order to allow analysis of patients with normal or failed karyotypes.38  Such studies have an important role in early follow-up since they have the advantage of being quantifiable. The use of FISH for the detection of rare events is, however, limited by the background noise and by the number of nuclei that can realistically be examined. The increasing use of interphase FISH should be subjected to the same quality control as PCR analysis. In general, close collaboration between those undertaking PCR and FISH analyses is the best method of ensuring that the most appropriate technique is used in each clinical context.

Since Ig/TCR-associated translocations lead to abnormal expression of an essentially normal proto-oncogene, despite potential somatic mutations in the coding sequences, its detection by immunological techniques represents an alternative form of diagnostic analysis. This has the following advantages: it is independent of genetic heterogeneity; it represents the functional oncogenic endpoint and therefore potentially correlates more closely with prognosis; and it allows in situ immunohistological analysis. Its disadvantages include the inability to distinguish physiological from pathological expression and the absence of specific monoclonal antibodies that will work reliably on fixed material. Distinction of normal from pathological expressions of some oncogenes may not be a problem, if the wild type protein is not normally expressed in lymphoid tissue or is expressed with an easily identified architectural pattern in diagnostic material. For example, immunological detection of ALK expression in anaplastic lymphoma (see below) or, theoretically, BCL8 (Section III) in DLCL is useful since neither protein is expressed in normal hematopoietic tissues. Immunological detection of BCL2 is useful at diagnosis of follicular lymphoma, since the pattern of expression in the germinal centers allows distinction from reactive nodes. Use of BCL2 antibodies cannot, however, be used for staging or follow-up. In contrast, antibodies to BCL6 have little place in diagnostic assessment since this protein is expressed in normal germinal center B cells, in lymphomas with BCL6 rearrangements and/or mutations and in those with a germline configuration.39  Similarly, detection of PAX5 protein expression is unlikely to have diagnostic significance, since it is normally expressed in the B lymphoid lineage (Section III).

Clinical applications at diagnosis

Oncogenic markers are likely to play an increasingly important role in the classification of NHL, but it is not yet clear which techniques(s) will be most appropriate for their detection in a routine diagnostic setting. Several abnormalities often co-exist, particularly at later stages of the disease. Since cytogenetic abnormalities are often complex and the molecular breakpoints heterogeneous, identification of a given genetic marker has diagnostic/prognostic significance but failure to do so does not.

For example, detection of BCL2-IgH [t(14;18)] in follicular lymphoma by karyotype analysis at diagnosis is approximately 75% for involved lymph nodes and 50-70% by PCR using MBR and MCR primers, the discrepancy being largely due to variant breakpoints.40  Aberrant follicular expression of the BCL2 protein is found in approximately 90% of cases. Consequently, absence of BCL2 protein is therefore a more appropriate basis for calling into question a diagnosis of follicular NHL than absence of BCL2 DNA rearrangement. Further analysis of BCL2 negative cases will determine whether they represent functionally equivalent variants expressing, for example, other BCL2 family members or NHL with distinct pathogenic mechanisms.

Genomic breakpoints in MYC-IgH in Burkitt's lymphoma with t(8;14)(q24;q32) are much more heterogeneous, extending from over 300 kbp either side of the c-MYC coding region and frequently involving the IgH switch regions. They are therefore not readily detectable by PCR, and Southern blot analysis is only informative if positive. c-MYC-specific antibodies have limited practical application, and FISH analysis currently represents the most rapid and appropriate method for diagnosis.34 

The t(11;14) represents an intermediate situation, with approximately 40% of cases of MCL demonstrating breakpoints in the major translocation cluster (MTC) of the BCL1 region situated approximately 100 kb upstream to CCND1. As such they are easy to detect by PCR.41  The majority of cases in MCL, and virtually all those in B lymphoproliferative disorders other than MCL (E.A. Macintyre, unpublished observations and 42 ), however, do not involve this. MTC and their breakpoints are best detected by FISH, preferably by strategies that also work on interphase cells.43  For diagnostic purposes, it is clear that detection of CCND1 expression, the functional endpoint of these genetic abnormalities, is the most appropriate tool. This can either be performed by analysis of RNA expression44,45  or by immunohistochemistry.46  Interpretation of the diagnostic significance of such expression cannot, however, be undertaken in isolation, since CCND1 is seen in a wide variety of B lymphoproliferative disorders with different prognoses and therapeutic requirements, including MCL, multiple myeloma, hairy cell leukemia, prolymphocytic leukemia and splenic lymphoma with villous lymphocytes.42,47,48  In practice, we restrict BCL1-IgH PCR to analysis of CCND1 RT-PCR-positive cases and refer CCND1-positive, BCL1-IgH-negative cases to further molecular cytogenetic analysis by fiber FISH49  in an attempt to determine whether other genetic clusters exist and whether these vary with diagnostic subtype.

The NPM-ALK fusion transcript in anaplastic large cell lymphoma (ALCL) illustrates nicely the way in which diagnostic practice evolves rapidly. The demonstration that certain CD30+ ALCL are associated with a t(2;5)(p23;q35) that involves NPM and ALK allowed diagnosis by RT-PCR50  and, in some cases, by DNA PCR.51  Limited availability of material suitable for RNA analysis restricted application of the former. With the development of ALK-specific monoclonal antibodies,52  it became clear that not only did the vast majority of NPM-ALK-positive cases aberrantly express ALK, but that ALK-positive, NPM-ALK-negative cases exist.53  Furthermore, the localization (nuclear and/or cytoplasmic) of ALK allowed distinction of certain variants.54  At the present time, ALK immunophenotypic analysis probably represents the most appropriate technique at diagnosis, with RT-PCR being reserved for confirmation, partner identification and initial characterization if follow-up is planned.

Prognostic evaluation requires assessment of a large number of genetic targets. As detailed in Section III, the number of recognized oncogenes involved in B cell lymphomas is increasing rapidly. Deregulation of distinct oncogenes, such as BCL2 and BCL6 in follicular lymphoma, can coexist at diagnosis and is even more common at relapse. At least two IgH switch translocations in myeloma have been shown to simultaneously deregulate expression of two independent oncogenes from both the chromosome 14 and partner derivative chromosomes by juxtaposing them to the Eα and Eμ IgH enhancers, respectively: FGFR3 and MMSET/WHSC1 in the t(4;14)(p16;q32)55,56  and myeov and CCND1 in the t(11;14)(q13;q32).57  It is probable that this will also be the case for NHL switch translocations. Prognostic assessment also depends on evaluation of the direct or indirect consequences of genetic abnormalities, which are unlikely to be the same for all patients in a given diagnostic category. For all these reasons, specific detection of individual abnormalities is increasingly unrealistic, and it is likely that global assessment of gene expression in lymphoma by microarray technology58,59  will replace detection of specific abnormalities (see below).

Practical Aspects

Improvements in our understanding of the mechanisms and consequences of genetic abnormalities in NHL will only improve clinical management if appropriate pathological material is available for the majority of patients. Certain practical aspects of analysis of lymphoma biopsies render the generation of high-quality nucleic acid preparations more problematic than for their leukemic counterparts, particularly for RNA and high molecular weight DNA. Biopsies are performed, frequently at a distance from the pathology laboratory, by surgical personnel who are often less conscious than their hematologyoncology colleagues of the importance of appropriate conservation conditions. While it is possible to obtain PCR-grade DNA suitable for the amplification of small fragments (up to approximately 500 bp in length) from appropriately fixed material, only fresh or cryopreserved material is suitable for cytogenetic, RNA-based PCR and most Southern blot analyses. Furthermore, variations in fixation conditions (type of fixative, use of buffered solutions, duration of fixation) have a significant effect on amplification of even small DNA fragments.60,61  Given the increasing number of (cyto)genetic analyses that can only be performed from fresh or cryopreserved material, it is our opinion that the priority for molecular diagnostic laboratories should be to encourage the establishment of surgical and laboratory procedures that allow conservation of unfixed material in all cases. This implies centralizing lymphoma biopsies in locations where fresh material can be rapidly transferred to a laboratory, where it can be sub-divided into fractions destined for cyto/histological, cytogenetic, immunological and molecular analysis, based on locally defined priorities for cases with limited material. Alternatively, the creation of specialized hemtological malignancy centers that regroup the aforementioned expertise will optimize rational use of material, in addition to allowing an intergrated approach to patient management.62  While it is obviously desirable to be able to analyze fixed material, attempts to adapt the technology to its systematic use will retard the evolution of the place of molecular analyses in the diagnosis and classification of lymphoid disorders.

The heterogeneous nature of lymphomatous infiltration can complicate interpretation of results. Whenever possible, it is preferable to conserve sections that correspond to the material undergoing molecular analysis for histological assessment in case of discrepant results. The interpretation of molecular analyses of tissues with minimal morphological involvement is always more reliable, and often only possible, when baseline analysis of pathological diagnostic material is available. This is particularly true when Ig/TCR clonal markers are used.

PCR-based techniques also suffer from their sensitivity. Increased recognition of this problem has led to a diminution in the incidence of contamination due to the establishment of appropriate laboratory structure and techniques and to the inclusion of control samples. Despite this, a recent international assessment of PCR detection of BCL2-IgH rearrangements showed a 28% false-positive rate among laboratories with records of publication in molecular diagnostics.63  The risk of contamination obviously correlates with the effort expended to detect rare events, and all attempts to minimize the use of nested PCR and the manipulation of PCR products, such as with the use of quantitative real time PCR (RQ-PCR) analyzers, will decrease this risk. In general, if a single round PCR reaction is insufficient for detection of a given target at diagnosis in fresh or cryopreserved material in which the malignant population represents at least 10%, either it is present in a minor sub-population and may not be representative of the tumor in general or the PCR should be optimized. This should not represent an indication for nested PCR. The identification of very low levels of an increasing number of molecular oncogenic markers such as BCL2-IgH64,65  in normal individuals also complicates interpretation of low level positivity. It has been our practice to limit nested analysis to follow-up of known molecular markers and to avoid making molecular diagnoses on uninvolved material destined more appropriately for staging/follow-up analysis. The increasing use of RQ-PCR is obviously reducing the indications for nested strategies.

Detection of contamination is more difficult for fusion transcripts since the PCR products from all individuals are identical, whereas V(D)J rearrangements are clone specific and consequently highly variable. Conversely, since the vast majority of contaminants are PCR products, repeated analysis of RNA in the absence of reverse transcriptase will permit their detection, whereas this is not possible for DNA PCR. Hybridization of transferred PCR products increases specificity but will not distinguish false from true positive results. An alternative strategy involves performing a confirmatory, “shifted” one-round PCR from RNA or DNA using a second pair of oligonucleotides, one of which is external to the oligonucleotides used for initial screening (and used only for this purpose in order to limit the risk of contamination).66  If appropriately sized products are obtained in both PCR, the specificity is provided by four hybridizing sequences, rather than three with hybridization strategies, and potential contaminants are unlikely to be positive with both PCR.

Despite these limitations, molecular analytical techniques have several advantages, including rapidity, simplicity, and economy. Retrospective analysis of appropriately archived material is possible. The small quantities of DNA/RNA required for PCR analysis (0.1-1 μg) makes analysis of material such as endoscopic or skin biopsies and cerebrospinal fluid feasible.24,25  With increasing experience and technological standardization, molecular diagnostics are likely to find an increasingly important place in lymphoma diagnostics. An ongoing BIOMED 2 European Concerted Action program, “PCR-based clonality studies for early diagnosis of lymphoproliferative disorders” coordinated by J.J.M. van Dongen, represents an attempt to assist this process.

Detection of Minimal Residual Disease

Space constraints prevent a detailed review of techniques for detection of minimal residual disease and only general comments will be made here. In the early days following the application of PCR analysis to detection of rare events in hematological malignancies it was thought that maximal sensitivity would be optimal, that it would allow analysis of the “black hole” represented by complete morphological remission and that it would translate into direct clinical benefit. With 15 years of experience, it is clear that not all PCR are equal and that the sensitivity required depends on the clinical situation. The role of molecular follow-up in individual patient management should also take into account that the majority of therapeutic decisions are taken at a relatively early stage and that identification of patients at high risk of relapse at this stage is of more potential benefit than the early prediction of relapse. Identification of high-risk patients is based on either the intrinsic characteristics of the tumor at diagnosis or the kinetics of tumor response to treatment. Molecular assessment of the latter has been shown to represent a powerful, independent prognostic marker in childhood ALL, based on Ig/TCR clonality analysis.10,11 

The situation is, however, more complicated in NHL. Heterogeneous involvement of blood or bone marrow in all but leukemic forms of NHL leads to sampling inaccuracies. Molecular genetic follow-up of fusion transcripts has the advantage of being comparable between patients as a result of target homogeneity. In contrast, the variable nature of Ig containing translocations renders some targets more readily amplifiable than others. This is also the case for lymphoid clonality. These strategies are obviously applicable only to marker-positive patients, but molecular follow-up can have significant clinical impact only if applicable to the majority of patients in a therapeutic protocol. At the moment, this is the case only for BCL2-IgH in FL, NPM-ALK in ALCL and IgH in somatically unmutated NHL, such as MCL. Pooling of results obtained with different targets was not possible with qualitative PCR strategies and was questionable with semi-quantitative, competitive strategies. It is to be hoped that RQ-PCR techniques will allow the generation of quantitative follow-up irrespective of the marker employed, but this remains to be demonstrated in NHL.

Evidence that molecular techniques can be used for practical patient management is largely based on BCL2-IgH and, to a lesser extent BCL1-IgH and lymphoid clonality.67,68,69,70,71 

Future Perspectives and Conclusions

The imminent achievement of the human genome sequencing projects and the arrival of microarray technologies58  are revolutionizing biological evaluation of NHL. Gene expression profiling has been shown to allow stratification of DLBCL patients into those with a germinal center-like profile and a relatively good prognosis when compared with those demonstrating an activated B lymphocyte-like profile.59  Our current collective priority is to find the best way to rapidly and efficiently evaluate the use of these techniques both to direct day to day patient management and to improve our understanding of lymphomagenesis. We are increasingly being faced with an excess of genetic data. The role of biologists interested in lymphoma management will be to order this information with respect to current biological practise. Cooperation at a national or international level is an inherent part of this process, as is the exploitation of high quality, well characterised tissue banks. This in turn raises economic, structural, educational, ethical and legal issues. National scientific and health care structures are responding to this “sea change” in biological practice in different ways.

The rapid increase in the incidence of NHL, in the rate at which novel genetic markers are being identified and in our understanding of the mechanisms leading to these abnormalities makes for an exciting period in biological lymphoma management. It is to be hoped that these developments will provide the basis for novel therapeutic strategies That will translate into improved clinical outcome.

Dennis Willerford, M.D.*

*

Division of Hematology, University of Washington, Box 357710, Seattle WA 98195-7710

Chromosomal translocations involving antigen receptor genes are a central feature of many lymphoid neoplasms, including the majority of NHLs as well as some acute leukemias and plasma cell neoplasms.1,2,3,4  These translocations characteristically juxtapose a cellular gene regulating cell growth, survival, or differentiation with transcriptional enhancer elements for antigen receptor genes, leading to deregulated oncogene expression. While other genetic changes are required for malignant transformation, in many cases the translocation is sufficient to initiate the tumor pathway, suggesting that it represents an early and essential step in lymphoid oncogenesis. Two components can therefore be considered in the pathogenesis of lymphoid malignancies: 1) generation of potentially oncogenic translocations; and 2) evolution of clinical malignancy as a result of growth or survival advantages and further genetic changes. The latter component is likely to be polymorphous, reflecting individual pathways for different oncogenes—corresponding in turn to distinct clinical entities.4  In contrast, the recurrent involvement of antigen receptor genes, particularly the immunoglobulin (Ig) heavy-chain locus, suggests a restricted set of mechanisms for creation of antigen receptor translocations. As a consequence, it could be that a large fraction of lymphoid malignancies is accounted for by a limited set of genetic, environmental and immunological circumstances contributing to translocation events.

Lymphoid cells utilize several unique forms of physiological genomic instability in order to create the vast diversity of Ig and T cell receptor (TCR) genes which recognize antigens. V(D)J recombination assembles Ig and TCR variable region gene segments during lymphoid development and during immune responses. In addition, immunoglobulin genes undergo isotype switch recombination and somatic hypermutation in antigen-stimulated B cells. The correlation between these unique physiological forms of genomic instability and the recurrent involvement of antigen receptor genes in oncogenic translocations in lymphoid malignancies suggests that lymphoid cancers may arise from errors in physiologic recombination events. Although direct evidence for this hypothesis and mechanistic insights remain limited, recent advances in three areas provide new opportunities for dissecting the underlying causes of lymphoid cancers: identification of genes involved in physiologic antigen receptor recombination and progress in understanding both the mechanisms of DNA cleavage and rejoining; improved understanding of how genomic instability is regulated in antigen-stimulated B cells; and molecular and genetic studies of human disorders and gene-deficient mice characterized by susceptibility to antigen receptor translocations and defective cellular responses to DNA damage.

Somatic Rearrangement and Mutation of Antigen Receptor Genes

V(D)J recombination

Antigen receptor diversity is generated by DNA rearrangement of germline variable (V), diversity (D) and joining (J) segments (or V and J segments) to form a contiguous variable region exon (reviewed in 5,6 ). Combinatorial diversity created by selection among multiple germline segments, along with extreme sequence diversity created at the point of assembly, leads to a vast repertoire of antigen specificities and forms the basis for adaptive immunity.4,5  V(D)J recombination begins with site-specific double-stranded DNA scission that is targeted by characteristic recognition signal sequences (RSS) flanking the borders of recombining variable region gene segments. DNA cleavage is coordinated between two rearranging segments and is catalyzed by the Rag-1/Rag-2 endonuclease complex, leaving a blunt RSS (signal) end and a sealed hairpin structure at the coding end.7,8  Absence of Rag-1 or Rag-2 blocks lymphoid development beyond an early progenitor stage in mice and causes a subset of autosomally inherited severe combined immunodeficiency in humans.4,9 

The distinct structures of signal and coding ends created by Rag-mediated DNA cleavage are processed by different pathways.4,6  Coding end processing injects a measure of randomness into the primary sequence at the resulting joint. The sealed hairpin structure is opened, often asymmetrically, and frequently loses bases to nuclease activity. Additionally, expression of terminal deoxynuclotidyl transferase (TdT) in developing lymphocytes adds bases (N-nucleotides) in a template-independent fashion to the free ends prior to ligation. The extreme sequence diversity at the site of V(D)J coding joints corresponds to the antigen contact residues of TCR and Ig molecules and contributes to the diversity of the immune repertoire. It is this junctional diversity that is exploited for PCR detection of lymphoid clonality (Section I). Rejoining of coding ends requires the DNA-dependent protein kinase (DNA-PK), which is mutated in the mouse Scid strain, along with components of the nonhomologous end joining (NHEJ) pathway for DNA repair. The latter includes the Ku70 and Ku86 nuclear antigens, along with XRCC4 and mammalian DNA ligase IV.4  In contrast, signal ends are usually religated without gain or loss of germline sequence, leading to an excision circle or a chromosomal inversion event, depending on the orientation of the recombining segments. Resolution of signal ends requires the NHEJ components, but is accurate in the absence of DNA-PKcs. A summary of the genes involved in V(D)J recombination is provided in Table 1. The V(D)J recombination mechanism is shared among seven rearranging antigen receptor genes in T and B cells. Inasmuch as these genes rearrange differently in each lineage and at specific stages, additional regulatory mechanisms are required to target the correct loci for rearrangement. This involves remodeling of chromatin structure, a process that is linked to transcription of the unrearranged antigen receptor locus, which invariably precedes rearrangement.10 

Table 1.

Genes involved in V(D)J recombination.

Gene
Properties
Role in V(D)J Recombination
Other Roles
Rag-1   Endonuclease   Initiation   None known  
Rag-2   Endonuclease   Initiation   None known  
DNA-PKcs   Protein kinase   Rejoining coding ends   Switch recombination  
Ku70   DNA end-binding;   Rejoining coding and signal ends   NHEJ pathway; Growth  
Ku80   DNA end-binding; TdT recruitment   Rejoining coding and signal ends   NHEJ pathway; Growth  
XRCC4   Activates DNA ligase   Rejoining coding and signal ends   NHEJ pathway; Brain development  
DNA Ligase IV   End joining   Rejoining coding and signal ends   NHEJ pathway; Brain development  
TdT   Nucleotide addition   Diversifies coding joins   None known  
Gene
Properties
Role in V(D)J Recombination
Other Roles
Rag-1   Endonuclease   Initiation   None known  
Rag-2   Endonuclease   Initiation   None known  
DNA-PKcs   Protein kinase   Rejoining coding ends   Switch recombination  
Ku70   DNA end-binding;   Rejoining coding and signal ends   NHEJ pathway; Growth  
Ku80   DNA end-binding; TdT recruitment   Rejoining coding and signal ends   NHEJ pathway; Growth  
XRCC4   Activates DNA ligase   Rejoining coding and signal ends   NHEJ pathway; Brain development  
DNA Ligase IV   End joining   Rejoining coding and signal ends   NHEJ pathway; Brain development  
TdT   Nucleotide addition   Diversifies coding joins   None known  

A role for V(D)J recombinase in creating oncogenic chromosome translocations has been proposed, based on comparison of translocation breakpoint sequences with normal V(D)J joints (reviewed in 1,4 ). Bearing among the strongest similarities to normal V(D)J joins are t(14;18) in follicular B cell lymphomas, where chromosome 14 breakpoints frequently occur at or near the RSS bordering DH or JH segments.1,11  Sequences resembling RSS have also been reported in 18q21 and other partner chromosomes. However, in many instances these bear poor homology to physiologic RSS or are located too far away from the breakpoint to play a role in Rag-mediated DNA cleavage.1,4  A novel translocation mechanism has been proposed based on the observation that the Rag proteins exhibit transposase activity in vitro; however, the extent to which this activity operates in vivo is unknown.12,13  Taken together, the available breakpoint sequence data suggest that most antigen receptor chromosome translocations do not represent the products of a normal V(D)J reaction. Nevertheless, it remains likely that unresolved DNA breaks in antigen receptor loci created during V(DJ) recombination are a factor in the translocation event.

Class switch recombination

Isotype class switching occurs in B cells stimulated by antigen and is accomplished by a deletional recombination event, replacing the Igμ/Igδ constant region gene cluster downstream of the functionally rearranged Ig variable region with a constant region encoding one of the Igγ, Igα or Igϵ isotypes. This leads to antibody molecules with identical antigen specificity but with specialized effector functions determined by the incoming constant region (reviewed in14 ). Recombination is targeted by repetitive switch region sequences located 5′ of Cμ and each of the constant region clusters, excepting Cδ. Class switching is regulated by T cell signals including cytokines as well as CD40 ligand/CD40 binding, which act in part by regulating germline transcription from noncoding exons located 5′ of the constant region clusters.14,15  Germline transcripts in switch recombination appear to play a direct role in initiating DNA breaks, through interaction with switch regions to produce DNA/RNA hybrid structures, which attract nuclease activity.16  Rejoining of DNA ends in switch recombination occurs in a reciprocal fashion17  and requires DNA-PKcs as well as the downstream components of the NHEJ pathway.18,19,20  Thus, V(D)J and class switch recombination have distinct mechanisms for creating double-strand DNA cleavage but appear to share a common pathway for resolution of these breaks. Ig class switch recombination is implicated in t(8;14) in some cases of sporadic Burkitt's lymphoma, where breakpoints occur within the switch regions upstream of Cμ, Cγ, or Cα,21,22  as well as in murine plasmacytomas, where c-myc is juxtaposed with the switch a region on mouse chromosome 12.23  Recently, a novel oncogenic mechanism based on switch recombination has been identified in a multiple myeloma cell line. An excised switch region product containing Cα and the 3′ IgH enhancer were found inserted into chromosome 11 in close proximity to the cyclin D1 proto-oncogene, associated with cyclin D1 overexpression.24 

Somatic hypermutation

During immune responses, germinal center B cells undergo further diversification of Ig specificities by the induction of somatic hypermutation in the V regions of IgH and IgL genes (reviewed in25 ). Subsequent selection processes favor survival and expansion of cells with increased affinity for antigen. Mutations occur at a rate of approximately 1/1000 bp per generation, and are confined to a region within about 1.5 kb downstream of the promoter. Extensive studies using transgenic mice indicate that transcription and somatic hypermutation are intimately linked and likely involve DNA repair processes, although the precise mechanism is not understood25,26 ). Somatic hypermutation most commonly involves base substitutions; however, recent studies have indicated that insertions or deletions may also occur with significant frequency, implying that the mechanism may involve DNA strand breaks.27,28,29  There is evidence that some Ig locus translocations may arise during the hypermutation process. For example, in some t(8;14) in Burkitt's lymphomas, the translocation involves a V(D)J rearranged IgH allele, and there are several examples of breakpoints occurring within the V region, correlating with the physiologic target of somatic hypermutation.27,30  Somatic hypermutation may also introduce genetic alterations outside of the Ig loci. When placed near Ig loci by a translocation, the c-myc and bcl-6 genes frequently develop point mutations characteristic of somatic hypermutation, potentially contributing to the additional genetic changes required for progression of the malignant phenotype.31,32  Recently, it was shown that somatic hypermutation also affects the germline bcl-6 gene in normal germinal center B cells, indicating that this process is not completely restricted to the Ig loci under physiologic circumstances.33,34 

Genomic Instability in Germinal Center B Cells

Antigen receptor diversification in lymphocytes occurs not only during development in the primary lymphoid organs but also as part of immune responses. This is particularly the case with B cells, which undergo a second phase of development in germinal centers. Germinal centers are specialized lymphoid organs that arise from a limited number of antigen-activated B cells, which migrate to primary lymphoid follicles and interact with antigen-bearing follicular dendritic cells (FDC).35,36  There is intense proliferation of the oligoclonal B cell population, where cycling centroblasts segregate from quiescent centrocytes in discrete dark and light zones, respectively. In the dark zone, proliferating B cells undergo somatic hypermutation. B cell progeny bearing mutated Ig genes may exit the cell cycle and migrate to the light zone, where they interact with FDC bearing antigen and antigen-specific T cells. These interactions promote the survival of cells with high-affinity antigen receptors, leading to the rapid evolution of humoral immunity toward high-specificity effector mechanisms. Interactions between centrocytes and T cells also facilitate Ig class switching.37  Centrocytes selected by antigen may re-enter the dark zone and undergo further clonal expansion and somatic hypermutation, or may exit the germinal center to differentiate into memory B cells or plasma cells.

V(D)J recombination of antigen receptor genes has until recently been regarded as an ordered differentiation process. In recent years it has become increasingly apparent that productively rearranged antigen receptor genes may undergo additional rounds of V(D)J rearrangement, a process termed receptor revision.38  In the Ig light chain genes, this may involve joining of previously unrearranged V and J elements, while in the IgH locus, V-region replacement may occur via a cryptic RSS element present near the 5′ end of the coding regions in the majority of V segments. In immature IgM+ bone marrow B cells, receptor revision occurs in response to Ig receptor signals, leading to a change in specificity, which may assist in preventing the emergence of autoreactive Ig molecules.38,39  Rag-1 and Rag-2 are also expressed in peripheral B cells; such cells are markedly increased after antigenic stimulation, suggesting that receptor revision occurs during immune responses.40,41,42,43  The developmental stage of Rag-expressing peripheral B cells and the role of receptor revision in immunity are the subjects of intense interest at present. Somatic hypermutation and receptor revision by V(D)J recombination have also been identified in peripheral T cells; however, their role in regulating the peripheral T cell repertoire is not yet known.25,44 

Given the intense genomic instability that accompanies the germinal center reaction, it is plausible to hypothesize that some or all of the critical genetic alterations leading to malignant transformation may arise during immune responses. Nodal B cell NHL subtypes resemble normal components of the germinal center, and successive pathological classification schemes have built on improving knowledge of the presumed normal cellular counterparts.45  This connection is supported by the finding that many of these tumors harbor somatic mutations of Ig V regions (reviewed in 3,27,30 ), suggesting that the target of critical transforming events was the product of a germinal center reaction. Several clinical observations also support a link between immunity and lymphoma. Helicobacter pylori infection is associated with gastric MALT lymphomas.46  The malignant proliferation often retains an antigen-driven component, such that remissions may be obtained after eradication of the organism with antibiotics. A subset of the AIDS-related lymphomas is not associated with profound immunodeficiency or EBV-mediated lymphoproliferation, and appears to correlate with chronic immune stimulation. Lymphoma Ig specificities for HIV proteins or autoantigens have been documented in some cases.47,48  Lymphomas may also occur in association with chronic hepatitis C, sarcoidosis and autoimmune diseases. The common thread here would appear to be chronic immune stimulation, perhaps in the setting of disordered immune responses.

Genetic Susceptibility to Antigen Receptor Locus Translocations

Given the vigor with which Ig genes are modified both during B cell development and immune responses, it is perhaps a wonder that lymphoid malignancies are not more common. It seems likely that tumor suppressor mechanisms are active in lymphocytes to ensure that physiologic DNA breaks are not rejoined in an inappropriate fashion. Clues to these functions may be obtained through study of humans and mice with genetic alterations wherein the frequency of oncogenic antigen receptor translocations is increased (Table 2).

Table 2.

Gene deficiencies that confer susceptibility to antigen receptor translocations in lymphoma.

Gene
Function
Species
Lymphoma Type
Translocation
ATM   DNA damage response   Human   T-NHL; T-PLL   t(7;14)  
ATM   DNA damage response   Mice   Thymic   Multiple  
NBS1   DNA damage response   Human   B-NHL; T-NHL   Unknown  
BLM   Chromosomal maintenance   Human   NHL   Unknown  
*p53   DNA damage response   Mice   Pro-B   t(12;15) IgH-cmyc  
**DNA-PKcs   Rejoin V(D)J coding ends   Mice   Pro-B   t(12;15) IgH-cmyc t(12;12) IgH-nmyc  
**XRCC4   NHEJ pathway   Mice   Pro-B   t(12;15) IgH-cmyc  
**Ku86   NHEJ pathway   Mice   Pro-B   t(12;15) IgH-cmyc  
Gene
Function
Species
Lymphoma Type
Translocation
ATM   DNA damage response   Human   T-NHL; T-PLL   t(7;14)  
ATM   DNA damage response   Mice   Thymic   Multiple  
NBS1   DNA damage response   Human   B-NHL; T-NHL   Unknown  
BLM   Chromosomal maintenance   Human   NHL   Unknown  
*p53   DNA damage response   Mice   Pro-B   t(12;15) IgH-cmyc  
**DNA-PKcs   Rejoin V(D)J coding ends   Mice   Pro-B   t(12;15) IgH-cmyc t(12;12) IgH-nmyc  
**XRCC4   NHEJ pathway   Mice   Pro-B   t(12;15) IgH-cmyc  
**Ku86   NHEJ pathway   Mice   Pro-B   t(12;15) IgH-cmyc  
*

in combination with deficiency of V(D)J rejoining components DNA-PKcs, Ku86, or XRCC4.

**

when combined with p53-null mutation.

Human genetic syndromes

Among inherited syndromes in humans, lymphoid malignancies are increased in two types of disorders: immune defects that impair host responses to EBV, and a subset of autosomal recessive disorders affecting DNA damage responses. The latter includes ataxia-telangiectasia (A-T), Nijmegen breakage syndrome (NBS), and Bloom's syndrome. A-T is characterized by progressive cerebellar ataxia, telangiectasias on sun-exposed surfaces, hypersensitivity to ionizing radiation, and cellular and humoral immunodeficiency (reviewed in 49 ). The gene mutated in A-T, ATM, encodes a member of the PI-3 kinase gene family that includes the DNA-PKcs.50  A-T fibroblasts exhibit general chromosomal instability, and nonmalignant translocations involving TCR genes are frequently found in blood lymphocytes.51,52  Cancer occurs in approximately 38% of A-T patients,49,52,53  of which 85% are lymphoid leukemias and lymphomas. The majority of these are of T cell origin and exhibit recurrent TCR locus translocations, often involving the TCL1 oncogene on chromosome 14q32.54  NBS shares several features in common with A-T, including hypersensitivity to radiation, the characteristic chromosomal rearrangements in lymphocytes involving antigen receptor loci, and the predisposition toward malignancies, particularly those of lymphoid origin (reviewed in 55 ). In contrast to the spectrum of malignancy in A-T, the most common lymphoid tumors in NBS patients are NHL of B cell origin. The NBS1 gene encodes nibrin, which forms a multi-protein complex that includes Rad50 and Mre11.56,57  Following exposure to radiation this complex forms nuclear foci, a function that is defective in nibrin-deficient cells.56  Bloom's syndrome is an autosomal recessive disorder characterized by defective growth, defects in humoral immunity, and an extraordinary incidence of cancer. Bloom's syndrome cells have a mutator phenotype and a distinct pattern of chromosomal instability, including recombinational exchanges between homologous chromosomes.4,58  Mutations are found in the BLM gene, which is homologous to ReqQ DNA helicases in bacteria. Cancer afflicts the majority of Bloom's syndrome patients, usually at a young age; the most common tumors are lymphomas and lymphoid leukemias.

At present, there is only a superficial understanding of how the ATM, nibrin and BLM proteins function in cells. Both A-T and NBS patients and respective cell lines exhibit heightened sensitivity to agents inducing double-strand DNA breaks, such as ionizing radiation or bleomycin, as well as abnormal cell-cycle regulation in response to DNA damage. Neither protein appears to be directly involved in repairing DNA damage.59  Similarly, a direct role for ATM, nibrin or BLM in V(D)J recombination has been discounted based on normal rearrangement of plasmid recombination substrates in patient fibroblasts,60  (C. Yeo, P. Concannon, and D.M. Willerford, manuscript in preparation). ATM has protein serine-threonine kinase activity, and a number of substrates have been identified in vitro and in vivo, many of which tend to play roles in either cell-cycle checkpoint control or apoptotic pathways.61,62,63  In particular, ATM phosphorylates p53 in response to DNA breaks, and accumulation of p53 in response to DNA damage is delayed or absent in A-T cells. Recently, it has been shown that ATM also phosphorylates nibrin.64,65  The emerging picture of ATM function is that it plays a signaling role, receiving information from molecules that detect DNA damage and relaying that information to appropriate downstream effector molecules that trigger responses ranging from cell cycle arrest to apoptosis. In this context, ATM and its downstream signaling partners could respond to unresolved DNA breaks during V(D)J recombination, either suppressing indiscriminate repair pathways, or perhaps limiting the survival of cells at increased risk for abnormal rejoining events.

Susceptibility to lymphoma in gene-deficient mice

Lymphomas arising spontaneously in genetically mutant strains of mice represent potentially important models for understanding the pathogenesis of lymphomas in humans. Recently, mouse strains exhibiting susceptibility to translocation-associated lymphoid malignancies have been identified, providing potential clues to the translocation process. In mice with combined Scid and p53-null mutations there is a high incidence of B lineage tumors occurring at an early age, compared with tumors exhibiting longer latency and predominantly thymic origin in the parental p53-/- strain.66,67,68,69,70  The majority of these tumors exhibit translocations involving the IgH locus on chromosome 12. Translocation partners include regions near c-myc on chromosome 15 and n-myc on chromosome 12, frequently accompanied by amplification of the translocated segment and associated with high expression of myc proteins.70  The tumors appear to arise at the pro-B cell stage, coincident with physiologic IgH rearrangement, suggesting that the translocations arise during attempted IgH locus rearrangement in Scid pro-B cells. This conclusion is supported by the finding that a Rag-2-null mutation blocked the development of t(12;15) pro-B cell lymphomas when introduced into the Scid p53-/- strain, demonstrating that initiation of V(D)J recombination was a required element in the oncogenic pathway.70  These studies suggest that physiologic suppression of oncogenic DNA rearrangements during V(D)J recombination has two important elements: efficient rejoining of DNA ends created by Rag-mediated DNA scission, and an intact cellular response to DNA damage. This model is supported by recent observations showing that mutations in other components in the V(D)J rejoining pathway, including Ku86 and XRCC4, lead to a similar phenotype of pro-B cell tumors bearing t(12;15) when bred onto the p53-deficient background.71,72,73  In the absence of effective DNA damage responses, accumulation of unresolved V(D)J ends may permit repair by alternative DNA mechanisms that are indiscriminate with regard to rejoining.

Conclusions

By pointing to oncogenes, chromosome translocations involving antigen receptor loci have provided critical insights into the process of cellular transformation in lymphoid malignancies. By comparison, little is known regarding how and why these translocation events arise. Direct evidence is accumulating that physiologic genomic instability aimed at diversifying antigen receptor genes predisposes lymphoid cells to this lineage-specific oncogenic mechanism, and it may soon be possible to achieve a general understanding of the causes of these tumors. This will require a more detailed understanding of the mechanisms of DNA breakage and rejoining in lymphoid cells, as well as of the genetic safety measures which suppress oncogenic translocations. Disruption of these mechanisms, whether by genetic variation, environmental exposures, or disordered immunity, may increase the likelihood of lymphomagenic translocations. The means to test these hypotheses directly is emerging, and may lead to strategies to manage individual and environmental risks.

Stephan W. Morris, M.D.*

*

Pathology, St. Jude Children's Research Hospital, 332 North Lauderdale, Thomas Tower, Room 5024, Memphis TN 38105-2729

Supported by National Cancer Institute (NCI) grants CA69129 and CA87064 (S.W.M.), CORE grant CA21765, and by the American Lebanese Syrian Associated Charities (ALSAC), St. Jude Children's Research Hospital.

Significant progress has been achieved over the past 20 years in the molecular characterization of various genetic abnormalities of pathogenic importance in NHL, spurred by the realization that characteristic chromosomal translocations, or rearrangements, tend to occur specifically in a given type of NHL, consistently altering the regulation of a particular gene. In the sections that follow, oncogenes involved by several of these recurrent chromosomal rearrangements in NHL are described. Due to space limitations, emphasis is placed on the more recently identified lymphomagenic genes; the interested reader will find a number of excellent reviews describing those genes identified in earlier studies to be important in the development of NHL (Table 3), such as the G1 cell cycle control gene cyclin D1 (CCND1, BCL1), the antiapopotosis gene BCL2, the bHLH/leucine zipper transcription factor cMYC, the IkappaB(BCL3)/REL/NFkappaB transcriptional pathway genes, and the POZ/zinc finger transcription factor gene BCL6 that are altered predominately in mantle cell, follicular, small noncleaved cell (Burkitt's), extranodal diffuse large-cell, or nodal diffuse large-cell NHL, respectively.1,2,3,4,5  Likewise, the role of aberrant anaplastic lymphoma kinase (ALK) gene expression in the genesis of NHL (“ALKomas”) will not be detailed here, given that B cell NHLs appear to be only rarely ALK-positive (almost all ALK lymphomas being T-lineage or null).6,7  The major characteristics of oncogenes deregulated in B cell NHL are summarized in Table 3.

Table 3.

Oncogenic genes deregulated by chromosomal translocations in B-cell non-Hodgkin's lymphoma.

Histology
Chromosomal Translocation
Activated Gene or Fusion Gene
Mechanism of Activation
Predominant Structural/Functional Features of Altered Gene(s)
Frequency
Follicular center cella  t(14; 18)(q32;q21)  BCL2  Relocation to IgH locus   Anti-apoptosis protein   80%  
Follicular center cell   t(1;22)(q22;q11)  FCGR2B  Relocation to IgL locus   ITIM-containing low affinity Ig Fc receptor   < 1%  
Small noncleaved cellb  t(8;14)(q24;q32)  cMYC  Relocation to IgH locus   bHLH/zip transcription factor   
  t(2;8)(p12;q24)  cMYC  Relocation to IgK locus   bHLH/zip transcription factor   100%  
  t(8;22)(q24;q11)  cMYC  Relocation to IgL locus   bHLH/zip transcription factor   
DLCLc  t(3;14)(q27;q32)  BCL6  Relocation to IgH locus   Zinc finger transcription factor   
  t(3;22)(q27;q11)  BCL6  Relocation to IgL locus   Zinc finger transcription factor   
  t(2;3)(p12;q27)  BCL6  Relocation to IgK locus   Zinc finger transcription factor   30-40%  
  t(3;4)(q27;p11)  BCL6  Relocation to TTF locusd  Zinc finger transcription factor   
  t(3;11)(q27;q23)  BCL6  Relocation to BOB1/OBF1 locusd  Zinc finger transcription factor   
DLCL   t(14;15)(q32;q11-13)  BCL8  Relocation to IgH locus   No homology with known proteins   <5%  
Mantle cell   t(11;14)(q13;q32)  CCND1 (cyclin D1)  Relocation to IgH locus   G1 cyclin   >95%  
Lymphoplasmacytoid   t(9;14)(p13;q32)  PAX5  Relocation to IgH locus   Paired homeobox transcription factor   50%  
MALT   t(11;18)(q21;q21)  AP12-MLT/MALT1  Gene fusion   Anti-apoptosis protein (API2), unknown (MLT/MALT1)   19-36%  
MALT   t(1;14)(p22;q32)  BCL10  Relocation to IgH locus   Apoptosis regulatory protein   <5%  
Splenic MZ lymphoma   t(2;7)(p12;q21)  CDK6  Relocation to IgK locus   Cell cycle kinase   <5%  
  t(7;14)(q21;q32)  CDK6  Relocation to IgH locus   Cell cycle kinase   
B-NHL   t(10;14)(q24;q32)  LYT10(NFKB2)  Relocation to IgH locus   REL transcription factor homology   <1%  
CLL/B-NHL   t(14;19)(q32;q13.1)  BCL3  Relocation to IgH locus   IκB   <1%  
B-NHL   ins(2;2)(p13;p11.2-14)  REL-NRG  Gene fusion   Transcription factor (REL), unknown (NRG)   <1%  
pre-B ALL (B-NHL ?)   t(5;14)(q31;q32)  IL3  Relocation to IgH locus   Growth factor   <1%  
pre-B ALL/B-NHL   t(1;14)(q21;q32)  BCL9  Relocation to IgH locus   No homology with known proteins   <1%  
B-NHL   t(1;14)(q21;q32)  MUC1  Relocation to IgH locus   Mucin-like glycoprotein   ∼5%  
Myeloma/B-NHL   t(1;14)(q21;q32)  MUM-2  Relocation to IgH locus   Fc receptor-related proteins  * 
  MUM-3    
B-NHL   t(11;14)(q23;q32)  DDX6/RCK  Relocation to IgH locus   DEAD box protein/RNA helicase   <1%  
B-NHL   t(12;22)(q13;q11)  CCND2(cyclin D2)  Relocation to IgL locus   G1 cyclin   <1%  
B-NHL   t(12;14)(q24;q32)  BCL7A  Relocation to IgH locus   Unknown function, weak homology to the actin-binding protein caldesmon   <1%  
The histologic spectrum of B-NHLs affected by most of the less frequent translocations has not been determined.  
Histology
Chromosomal Translocation
Activated Gene or Fusion Gene
Mechanism of Activation
Predominant Structural/Functional Features of Altered Gene(s)
Frequency
Follicular center cella  t(14; 18)(q32;q21)  BCL2  Relocation to IgH locus   Anti-apoptosis protein   80%  
Follicular center cell   t(1;22)(q22;q11)  FCGR2B  Relocation to IgL locus   ITIM-containing low affinity Ig Fc receptor   < 1%  
Small noncleaved cellb  t(8;14)(q24;q32)  cMYC  Relocation to IgH locus   bHLH/zip transcription factor   
  t(2;8)(p12;q24)  cMYC  Relocation to IgK locus   bHLH/zip transcription factor   100%  
  t(8;22)(q24;q11)  cMYC  Relocation to IgL locus   bHLH/zip transcription factor   
DLCLc  t(3;14)(q27;q32)  BCL6  Relocation to IgH locus   Zinc finger transcription factor   
  t(3;22)(q27;q11)  BCL6  Relocation to IgL locus   Zinc finger transcription factor   
  t(2;3)(p12;q27)  BCL6  Relocation to IgK locus   Zinc finger transcription factor   30-40%  
  t(3;4)(q27;p11)  BCL6  Relocation to TTF locusd  Zinc finger transcription factor   
  t(3;11)(q27;q23)  BCL6  Relocation to BOB1/OBF1 locusd  Zinc finger transcription factor   
DLCL   t(14;15)(q32;q11-13)  BCL8  Relocation to IgH locus   No homology with known proteins   <5%  
Mantle cell   t(11;14)(q13;q32)  CCND1 (cyclin D1)  Relocation to IgH locus   G1 cyclin   >95%  
Lymphoplasmacytoid   t(9;14)(p13;q32)  PAX5  Relocation to IgH locus   Paired homeobox transcription factor   50%  
MALT   t(11;18)(q21;q21)  AP12-MLT/MALT1  Gene fusion   Anti-apoptosis protein (API2), unknown (MLT/MALT1)   19-36%  
MALT   t(1;14)(p22;q32)  BCL10  Relocation to IgH locus   Apoptosis regulatory protein   <5%  
Splenic MZ lymphoma   t(2;7)(p12;q21)  CDK6  Relocation to IgK locus   Cell cycle kinase   <5%  
  t(7;14)(q21;q32)  CDK6  Relocation to IgH locus   Cell cycle kinase   
B-NHL   t(10;14)(q24;q32)  LYT10(NFKB2)  Relocation to IgH locus   REL transcription factor homology   <1%  
CLL/B-NHL   t(14;19)(q32;q13.1)  BCL3  Relocation to IgH locus   IκB   <1%  
B-NHL   ins(2;2)(p13;p11.2-14)  REL-NRG  Gene fusion   Transcription factor (REL), unknown (NRG)   <1%  
pre-B ALL (B-NHL ?)   t(5;14)(q31;q32)  IL3  Relocation to IgH locus   Growth factor   <1%  
pre-B ALL/B-NHL   t(1;14)(q21;q32)  BCL9  Relocation to IgH locus   No homology with known proteins   <1%  
B-NHL   t(1;14)(q21;q32)  MUC1  Relocation to IgH locus   Mucin-like glycoprotein   ∼5%  
Myeloma/B-NHL   t(1;14)(q21;q32)  MUM-2  Relocation to IgH locus   Fc receptor-related proteins  * 
  MUM-3    
B-NHL   t(11;14)(q23;q32)  DDX6/RCK  Relocation to IgH locus   DEAD box protein/RNA helicase   <1%  
B-NHL   t(12;22)(q13;q11)  CCND2(cyclin D2)  Relocation to IgL locus   G1 cyclin   <1%  
B-NHL   t(12;14)(q24;q32)  BCL7A  Relocation to IgH locus   Unknown function, weak homology to the actin-binding protein caldesmon   <1%  
The histologic spectrum of B-NHLs affected by most of the less frequent translocations has not been determined.  
a

BCL2 deregulation also occurs in 10-30% of DLCL and ∼30% of marginal zone lymphomas.

b

cMYC deregulation is also seen in ∼10% of DLCL.

c

Approximately 6-15% of follicular center cell and 10% of marginal zone lymphomas also have BCL6 deregulation.

d

Many other non-Ig loci have been reported also to deregulate BCL6.

*

Not yet determined

PAX5

Approximately 50% of small lymphocytic lymphomas with plasmacytoid differentiation, so-called lymphoplasmacytoid lymphoma (LPL), contain a t(9;14)(p13;q32). The t(9;14) illustrates the involvement in NHL of the PAX (for paired homeobox) transcription factor family that normally controls embryonic development and organogenesis.8,9  Members of this family contain two discrete DNA-binding domains—the paired box and the paired-type homeodomain—that display coordinate DNA binding specificity. Lymphomas containing t(9;14) possess a plasma cell-like phenotype with serum paraprotein production, and follow an indolent course followed by large-cell transformation. The t(9;14) results in juxtaposition of the PAX5 gene with the Ig heavy-chain gene on chromosome 14.10 PAX5 is normally expressed in fetal brain and liver during development but becomes restricted to B cells and testis after birth.11 PAX5 is transcribed throughout B cell ontogeny, but undergoes downregulation during plasma cell differentiation. Knockout mice experiments have demonstrated that Pax5 is important for midbrain development and that loss of Pax5 function results in maturation arrest of lymphocytes at the pro-B cell stage.12 Pax5 overexpression, by contrast, results in splenic B cell proliferation.13  Several genes important for B cell development have been proposed to be PAX5 targets including CD19, B cell receptor component Ig alpha (mb-1), transcription factors N-MYC and LEF-1 (positively regulated by PAX5), and the cell surface protein PD-1 and the p53 tumor suppressor (which are downregulated).14  Also, B cell SRC-family tyrosine kinase BLK, which transforms lymphoid progenitors in an activated form but is dispensable for B cell development and activation, is upregulated by PAX5.15,16,17  Further, PAX5 appears to be required for normal IgH VDJ recombination, given that V-to-DJ recombination is reduced ∼50-fold in Pax5-deficient pre-B cells.14 

Dysregulation of PAX5 transcription by a translocated IgH promoter has also been described rarely in NHL subtypes other than LPL and in myeloma.18  Interestingly, other PAX genes play a role in oncogenesis; for example, PAX3 and PAX7 fuse to the forkhead domain transcription factor (FKHR) in t(2;13) (q35;q14) and t(1;13)(p36;q14), respectively, in the skeletal muscle tumor rhabdomyosarcoma, and PAX5 itself is aberrantly expressed in some medulloblastomas and glioblastomas through an unknown mechanism.19 

BCL8

Translocations affecting chromosome 15q11-q13 and various partners occur in about 4% of diffuse large-cell lymphoma (DLCL). Recently, Dyomin et al identified the chromosome 15 target gene, BCL8, which is expressed as a major transcript of 2.6 kb and a less prominent 4.5-kb message due to differential polyadenylation.20 BCL8 expression is normally restricted primarily to the testis and prostate, with no transcripts found in hematopoietic tissues such as spleen, thymus, or blood leukocytes. RTPCR analysis of RNA from DLCL samples and lymphoma cell lines identified BCL8 expression in all cases having abnormalities of chromosome 15q11-q13, as well as in four of nine randomly selected DLCL cases and six of 15 DLCL cell lines (but none of three hyperplastic lymph nodes). The mechanism of aberrant BCL8 expression in lymphomas lacking 15q11-q13 rearrangements has not been defined.

The exact manner by which BCL8, which shares no homology with previously described genes, contributes to lymphomagenesis remains to be addressed. Likewise, the exact frequency of BCL8 abnormalities in NHL, whether aberrant BCL8 expression delineates a homogeneous clinical-pathologic entity, and the effect of BCL8 expression on the prognosis of NHL patients have yet to be investigated.

BCL9, MUC1, FcγRIIB and Other 1q21-q22 Genes

Abnormalities of the chromosome 1 long arm, in particular 1q21-q22, occur in a variety of preneoplastic and neoplastic lesions, in both solid and hematopoietic malignancies.21  The third most common rearrangement site in B cell NHL is band 1q21, after 14q32 and 18q21, the locations of IgH and BCL2, respectively. Abnormalities of 1q21-q22 occur in B cell malignancies arrested at all differentiation stages, and no clear association with a particular NHL subtype has been reported. Aberrations of this chromosomal region are usually secondary and occur in about 10-15% of B cell NHL, being associated with a poor prognosis, especially in DLCL.22  One of these 1q aberrations, the duplication dup(1)(q21q32), is the second most common cytogenetic abnormality in Burkitt's lymphomas (after cMYC translocation), occurring in more than 30% of cases. Breakpoints in 1q21-q22 show surprising heterogeneity and involve several target genes; in addition, there exists considerable promiscuity in the partners with which these 1q regions rearrange. It is important to point out that a number of 1q21 abnormalities result in an unbalanced chromosome 1 translocation, suggesting that in addition to activation of specific target genes, haploid loss of the chromosome 1 long arm may also contribute to oncogenesis in some tumors. As described below, progress has been made recently in the characterization of various 1q21-q22 target genes.

In 1998, Willis and colleagues reported the cloning of a t(1;14)(q21;q32) in a pre-B acute lymphoblastic leukemia cell line (CEMO-1), identifying the novel BCL9 gene as the 1q21 target.23 BCL9 encodes a 1,394-amino acid (aa) protein that contains several pentapeptide repeats, a potential nuclear localization signal, and a 30-aa region that shares 90% homology to a Drosophila embryo EST clone, but possesses no other recognizable domains. BCL9 is expressed ubiquitously as two low-level transcripts, a major 6.3-kb and a less prominent 4.2-kb message, while a 1.6-kb transcript is present only in spleen, thymus, and small intestine. The sequence of the 4.2- and 1.6-kb transcripts has not been reported.

The t(1;14) in CEMO-1 results in juxtaposition of IgH sequences adjacent to BCL9, the 1q21 break occurring in the 3′ untranslated region (UTR) of BCL9. Transcript levels of BCL9 in CEMO-1 are 50-fold higher than in EBV-transformed normal B cells. Southern hybridization screening of a panel of 3′ B cell malignancies containing 1q21 abnormalities showed only one additional case (an MCL) with a 30 UTR BCL9 breakpoint. FISH analysis using a BCL9-containing YAC identified one additional case (a follicular lymphoma) in which the gene was juxtaposed adjacent to the Igl locus from 22q11. The other 1q21 breaks in these 39 cases were heterogeneous, centromeric to BCL9 in half and telomeric in the remaining half. Thus, it was clear from this study that genes other than BCL9 are involved in many of the 1q21 rearrangements in B cell tumors.

Earlier this year, two reports described the characterization of a t(1;14)(q21;q32) in a large-cell lymphoma with an extranodal presentation (abdominal ascites), identifying dysregulation of the MUC1 gene, which is located ∼8 cM telomeric to BCL9.24,25  This translocation was shown to link IgHG4 switch sequences to the portion of chromosome 1 immediately downstream of MUC1, leaving MUC1 intact, and MUC1 transcription and translation were shown to be dramatically upregulated. None of six other genes located in an 85-kb region immediately centromeric to MUC1 (CLK2, propin, COTE1, GBA, metaxin, or thrombospondin-3) were found to be overexpressed in this tumor. Southern blot analysis of 72 B cell NHLs of all histologic types containing a 1q21 rearrangement revealed MUC1 rearrangement in 4 (6%) of the tumors (one case each of diffuse small cleaved, follicular mixed, immunoblastic, and extranodal large-cell). In addition, increased copy number of the MUC1 locus was identified in 18 (10%) of 178 B cell NHLs of all histologies, with the amplification being low level (4-6 copies).24 

MUC1 encodes a glycoprotein containing multiple copies of a tandemly repeated mucin-like domain. MUC1 mucin has been previously implicated in carcinogenesis. For example, adenocarcinomas with MUC1 mucin over-expression have high metastatic potential and poor prognosis; furthermore, Muc1null mice exhibit delayed progression of mammary tumors.26  MUC1 expression is also associated with inhibition of T cell proliferation, resulting in an “anergic” state and inhibition of cell-cell and cell-matrix interactions that facilitate metastatic tumor expansion.27  Expression of EMA (epithelial membrane antigen), which is equivalent to MUC1, occurs in essentially all ALCLs, and in lymphocyte-predominant Hodgkin's disease (75% of cases), plasmacytomas (75%), and T cell lymphomas (50%) due to unknown mechanisms other than 1q21 rearrangement. Thus, abnormal MUC1 expression may be involved in the progression of NHL, consistent with the involvement of 1q21 usually as a secondary abnormality in association with primary aberrations such as t(14;18)(BCL2) or t(8;14)(cMYC).

Yet additional 1q21-q22 gene loci undergo rearrangement in B cell NHL. Callanan et al recently reported analysis of a novel balanced t(1;22)(q22;q11) in three follicular lymphomas also containing t(14;18).28  This characterization revealed FCGR2B, which encodes the immunoreceptor tyrosine-based inhibition motif (ITIM)-containing low affinity IgG Fc receptor FcγRIIB, located at 1q22, as the target of this rearrangement. Three FcγRII genes (FCGR2A,-B, and -C) and two FcγRIIIB genes (FCGR3A and -B) are located in an ∼200-kb region in 1q22. The breakpoint characterized by Callanan et al interrupts the low affinity IgG Fc receptor locus just upstream of the FCGR3B promoter and 20 kb telomeric of the FCGR2B promoter, juxtaposing the 5′ region of FCGR2B next to the Igλ 3′ enhancer. At least 30-fold overexpression of the two major 1.5- and 2.4-kb FCGR2B transcripts was found in a t(1;22)(q22;q11)-positive follicular lymphoma cell line compared to other B cell lines such as Raji Burkitt's lymphoma or EBV-transformed lymphoblastoid cells; by contrast, FCG3RB expression was absent. RT-PCR analysis of t(1;22)-containing cases revealed FCGR2A and -C (which encode immunoreceptor tyrosine-based activation motif [ITAM] Fc receptors) to be expressed at levels equivalent to EBV-transformed B cells and control lines such as myeloid U937 cells. Very high levels of the FcγRIIB receptor alternative splice form FcγRIIb2 were specifically expressed in t(1;22)-positive cases, whereas b1 isoform levels were not elevated above normal. Thus, t(1;22)(q22;q11) results in constitutive, high-level expression of an intact FcγRIIB receptor (specifically, the b2 isoform).

FcγRIIB is one of the ITIM-containing inhibitory coreceptors that effect negative regulation of immune responses mediated by ITAM receptors such as B cell antigen receptors (BCR).29  For example, crosslinking of FcγRIIB to a BCR expressed on the same cell by IgG/ antigen complexes downregulates BCR-induced activation signaling due to recruitment of the inositol phosphatase SHIP to the complex by interaction with the FcγRIIB ITIM. It is presently unclear how high-level constitutive FcγRIIb2 expression might contribute to B cell lymphomagenesis, but Fcγ receptors can definitely affect B cell growth; for example, activation of these receptors by antibody binding enhances the growth and differentiation of murine B lineage progenitors in vitro, and FcγRII-deficient mice have a significantly increased B cell compartment.30  In addition, overexpression of FcγRIIB in nonlymphoid tumor cells can enhance tumorigenicity in vivo, perhaps by allowing the tumor cells to interfere with IgG activity, thereby diminishing antitumor antibody responses.31 

Cloning of yet another 1q21 breakpoint has revealed two additional Fc receptor family genes to be involved in the pathogenesis of B-lineage malignancies. Hatzivassiliou et al reported in abstract form the cloning of a t(1;14)(q21;q32) in a plasmacytoma cell line, identifying two genes within 20 kb of one another in the region spanning the 1q21 break.32  One of these genes, provisionally named MUM-2 (multiple myeloma-2), is expressed in spleen and lymph nodes as three transcripts of 2.6, 2.8, and 3.4 kb generated by alternative polyadenylation signals. The MUM-2 ORF predicts a 515-aa cell surface glycoprotein containing four extracellular Ig-type domains, a transmembrane and a cytoplasmic domain, with 37% identity (51% homology) to FcγRI over its first three extracellular domains. In the plasmacytoma from which it was cloned, the 1q21 breakpoint interrupts the MUM-2 coding region and juxtaposes it to IgH in the same transcriptional orientation, producing a fusion transcript between the first two MUM-2 exons and the transmembrane and cytoplasmic exons of Cα to encode a putative MUM2-Cα fusion protein. The second gene, MUM-3, is located 3′ to MUM-2, expresses three transcripts of 3.0, 5.0, and 6.5 kb in lymph nodes and spleen, and encodes a protein with an extracellular portion containing six Ig-type domains homologous to members of the Fcγ and Ig-type adhesion receptor families. MUM-3 expression is elevated not only in myelomas but also in Burkitt's lymphoma cell lines containing 1q21 abnormalities. The exact role that these novel cell surface receptors play in normal lymphocyte development and function, and in B cell malignancies, awaits further study.

API2-MLT/MALT1

Marginal zone lymphomas have been separated into three distinct disorders by the REAL classification—primary nodal, primary splenic, and extranodal lymphoma of the mucosa-associated lymphoid tissue (MALT) type.33  Although these disorders share a CD5- and CD10- negative B cell phenotype, they have subtle morphological differences and distinct clinical behaviors, suggesting that their pathogenesis may differ.34,35  Marginal zone lymphomas of the MALT are the most common subtype of lymphoma arising in extranodal sites, and account for 5-10% of all NHLs.36  These lymphomas have frequent multicentric extranodal involvement including the gastrointestinal tract, lung, thyroid, and mammary, salivary and lacrimal glands, and an indolent clinical course. Most MALT lymphomas originate in the setting of chronic inflammation triggered by infection or autoimmune disorders, including H. pylori gastritis, Sjögren's syndrome and Hashimoto's thyroiditis. Malignant B cell proliferation in gastric MALT tumors is dependent in part upon H. pylori-specific tumor-infiltrating T cells, and eradication of H. pylori by antibiotic treatment results in tumor regression in most patients.37  The t(11;18)(q21;q21) is a recurrent abnormality in marginal zone lymphomas, having been reported in up to 50% of extranodal low-grade MALT tumors. In 1999, Dierlamm and coworkers showed t(11;18) to produce a fusion of API2 at 11q21 with MLT (for MALT lymphoma-associated translocation) at 18q21.38  Simultaneously, two other groups also characterized the novel 18q21 gene, which both termed MALT1.39,40 

API2 belongs to a family of inhibitor of apoptosis proteins (IAPs) first identified in baculoviruses, in which they suppress host cell apoptotic responses to viral infection; five human IAPs have been identified—NIAP, API1 (also named cIAP1, HIAP2, MIHB), API2 (cIAP2, HIAP1, MIHC), XIAP-hILP, and survivin.41  IAPs contain a BIR (baculovirus inhibitor of apoptosis repeat) motif in one to three copies, a caspase recruitment domain (CARD), and a C-terminal zinc-binding RING finger domain (the latter being found in all IAPs except NIAP and survivin).42,43  IAP-1 and -2 were originally identified as proteins recruited to the tumor necrosis factor receptor II (TNFRII) cytoplasmic domain via association with the TNFR-associated factors (TRAFs), TRAF-1 and -2. API2 is highly expressed in lymphoid cells in the spleen and thymus and suppresses apoptosis by binding to and inhibiting caspases-3 and -7, as well as the cytochrome-c mediated activation of caspase-9.41  Engineered mutant IAPs that contain only BIR domains can bind and inhibit caspases, emphasizing the importance of these motifs.44 

The normal function of the novel MLT/MALT1 gene is unknown, but it encodes a protein homologous to a hypothetical C. elegans protein and contains two Ig-like C2-type domains similar to CD22β, a region with laminin 5 α3β homology, and a domain similar to mouse Ig-γ chain VDJ4.38,39,40 MLT/MALT1 is highly expressed in peripheral blood mononuclear cells, at moderate to weak levels in bone marrow, thymus and lymph nodes, and at high levels in hematopoietic cell lines of T-, B-, and myeloid lineages, suggesting a role in normal blood cell growth.

Although the exact contributions of the API2-MLT/MALT1 fusion to the genesis of MALT lymphoma remain to be determined, inhibition of apoptotic responses in MALT B cells is likely. Truncation of API2 C-terminal to its BIR domains may release their antiapoptotic effects from negative regulation mediated by the CARD and RING domains and, as mentioned, studies have shown that the BIR domain alone is sufficient for caspase inhibition and suppression of apoptosis,44  whereas over-expression of the RING domain alone increases cell death in a Drosophila system.45  The role of the C-terminal portion of MLT/MALT1 in the fusion is unclear, but all variants of the fusion identified (see below) have been inframe, supporting a specific function. A recent study of four API2-MLT/MALT1-positive cases revealed four-fold higher proliferation and survival in vitro as compared to t(11;18)-negative MALT lymphoma B cells.46 

RT-PCR of API2-MLT/MALT1 chimeric transcripts from a number of MALT tumors has revealed at least two breakpoints in API2 and three in MLT/MALT1, with at least four chimeric proteins predicted.47,48,49  In all cases, the N-terminus of the fusions contains all three API2 BIR domains, and the RING finger is eliminated (as noted, suggesting the fusion may liberate the BIR motifs from negative control by the RING finger). The presence of the API2 CARD is variable, but it is excluded in most of the fusions characterized; inclusion of the CARD could conceivably alter the caspase-binding abilities of the chimeric protein, given that a number of caspases also contain this homophilic interaction motif.42  The MLT/MALT1 sequence in the fusion varies significantly, the only identified motif invariably present being the C-terminal VDJ4-like domain. RT-PCR has revealed several differently-sized fusion products in individual tumors, suggesting that alternative MLT/MALT1 splicing (in addition to different breakpoints) may also contribute to variable API2-MLT/MALT1 transcripts.48  Expression of the reciprocal MLT/MALT1-API2 fusion transcript in only a subset of MALT lymphomas containing the t(11;18), together with the cryptic deletion of 3′ portions of API2 that would preclude expression of the reciprocal protein in other cases, suggest that the lymphomagenic properties of t(11;18) reside solely in API2-MLT/MALT1. Despite the heterogeneous breakpoints and complexity introduced by alternative MLT/MALT1 splicing, RT-PCR strategies can be designed to detect most, if not all, API2-MLT/MALT1 fusions. Interphase and metaphase FISH also seem to be robust methods for detecting t(11;18).38,39,40,50 

The clinical-pathologic characteristics of MALT lymphomas that express API2-MLT/MALT1 are still being defined, but preliminary comments can be made.47,48,49  For example, despite the origin from marginal zone B cells of primary nodal, primary splenic, and extranodal MALT-type marginal zone lymphomas, only the latter has been found to express API2-MLT/MALT, consistent with studies that have detected t(11;18) using other means such as FISH.50  Further, no evidence for t(11;18) has been found in any other type of NHL, indicating its apparent specificity. No clear-cut differences in the clinical features or survival of patients with API2-MLT/MALT1-positive MALT tumors as compared to t(11;18)-negative cases have yet been identified. However, two studies have identified a disproportionately high frequency of API2-MLT/MALT1-positive pulmonary MALT lymphomas,47,49  although t(11;18) is definitely present in MALT tumors occurring at gastric and other extranodal sites.48  MALT lymphomas harboring API2-MLT/MALT1 have not been shown to possess a greater likelihood of transformation to large-cell lymphoma, although the number of cases examined is relatively small.47,48,49  The incidence of API2-MLT/MALT1 detection in all extranodal lymphomas of MALT type, regardless of anatomic location, has ranged from 19-36% by RT-PCR analysis.47,48,49  For example, Baens et al identified 11 of 58 (19%) gastric MALT tumors to express API2-MLT/MALT1, and 48% of those gastric MALT tumors that lacked a large-cell component to be positive.48 

BCL10

The other recurrent translocation identified in MALT lymphomas, in addition to (11;18), is the t(1;14)(p22;q32),51  which is considerably less frequent. MALT lymphomas containing t(1;14) appear to be more clinically aggressive and grow more readily in vitro than those lacking the abnormality, and acquisition of t(1;14) is believed to contribute to development of H. pylori-independent gastric MALT tumor growth.

In 1999, Willis et al and Zhang and coworkers independently characterized t(1;14), identifying an apoptosis control gene, BCL10, as the 1p22 target.52,53 BCL10 is a CARD gene, like the 11q gene cIAP2 altered in t(11;18). The CARD is a homotypic interaction motif of ∼ 90 aa shared by proapoptotic (RAIDD/CRADD, CED4/Apaf1, caspases-1, -2, -4, -5, -9, -11, -12, and CED3, CARDIAK/RICK/RIP2) and antiapoptotic (cIAP1, cIAP2, ARC) molecules.42  CARD proteins are known to be essential for transducing death or survival signals; for example, CED3 and procaspase-9 associate with their regulatory proteins CED4 and Apaf-1, respectively, through CARD interactions that control activation of the caspase zymogens.54,55  Interestingly, despite its CARD, studies to date have not demonstrated BCL10 association via the motif with the CARD of any proteins other than itself.56,57,58,59 

The t(1;14) causes dramatic upregulation of BCL10 expression due to juxtaposition to the IgH enhancer.52,53  In normal tissues, BCL10 is ubiquitously expressed, although at relatively low abundance, as a 2.8-kb transcript. Tissues with highest BCL10 expression are spleen, lymph node, and testis. BCL10 is also expressed in B cell lines transformed at all stages of differentiation including pro-B ALL, pre-B ALL, Burkitt's lymphoma, multiple myeloma, and EBV-immortalized lymphoblastoid lines. The 233-aa BCL10 protein contains a single N-terminal CARD from residues 13 to 101 and a serine/threonine-rich C-terminus. Human BCL10 shares 29% identity (35.5% similarity) with ORF E10 of the gamma herpesvirus equine herpesvirus-2 (EHV-2).60  EHV-2 is restricted to horses, causing pharyngitis and lymphadenopathy; although not known to be lymphomagenic, EHV-2 is highly related to herpesvirus saimiri, which causes fulminant T-cell lymphomas in primates, and only slightly less related to EBV and human herpesvirus-8 that are implicated in Burkitt's lymphoma and lymphomas in immunocompromised patients.

Concomitant with the cloning of t(1;14), several apoptosis laboratories also published initial characterizations of BCL10 (referred to as cE10 [cellular E10], CIPER [CED3/ICH-1 prodomain homologous E-10-like regulator], CLAP [CARD-like apoptosis protein], CARMEN (CARD-containing molecule enhancing NFkappaB], or mE10 [mammalian E10]).56,57,58,59,61  These studies revealed enforced BCL10 expression in most cells (e.g., 293T, COS, HeLa, MCF7) to induce either modest56,61  or robust52,53,57,59  apoptotic death. A single study reported that BCL10 induces significant death only when coexpressed with death receptors (TNFR1, CD95/Fas), receptor-associated adapters (TRADD, FADD), or the receptor initiator caspase-8,56  while another failed to find significant effects upon receptor-mediated apoptosis.61  In variance with these reports of proapoptosis by BCL10 in many cells, stable BCL10 expression could be achieved in lymphoid cell lines (e.g., murine pro-B line BaF3 or BJAB Burkitt's lymphoma cells) without inducing an apparent propensity to apoptotic death.52,53 

All of the above-mentioned studies but one58  were in agreement that BCL10 expression activates NFκB. Dominant-negative forms of NIK, IKK, and IκB, but not dominant-negative TRAF2, TRAF6, or RIP, inhibit BCL 10-mediated NFκB induction, suggesting that BCL10 functions upstream of NIK, IKK and IκB, and either downstream or independent of TRAFs and RIP.56,57,61  Suggestive evidence for a direct role of BCL10 in TNFR-induced NFκB activation is the demonstration that normal BCL10 slightly enhances TNFα-mediated NFκB activation, whereas expression of the CARD domain alone inhibits activation.57  The exact manner in which BCL10 might couple to receptor NFκB activation pathways is not clear, however; for example, conflicting data exist regarding the ability of BCL10 to bind TRAFs1-6, one study reporting interactions with only TRAF1 and TRAF5,56  another identifying interaction with TRAF2,62  and others finding no interactions with any TRAFs.56,61  In addition, no interactions with NFκB signaling proteins CARDIAK/RICK/RIP2 or NIK have been identified.56,58  Despite a reported interaction of BCL10 with TRADD in one study, BCL10 mutants activated NFκB independent of TRADD association.61  Data from my laboratory are consistent with physical interaction of BCL10 and TRAF2, and suggest this interaction is important in regulating TRAF-2 mediated NFκB activation (S.W. Morris, unpublished data).

Given the data showing BCL10 to be proapoptotic in most cell types, overexpression of the gene due to t(1;14) appeared paradoxical. To address this issue, BCL10 transcripts from t(1;14)-positive MALT lymphomas were characterized and reported to contain a variety of mutations.52,53  Although no tight clustering of mutations was evident, these abnormalities tended to produce two distinct subsets of BCL10 mutants: (1) proteins of 33 to 101 aa truncated within the N-terminal CARD, and (2) proteins of 141 to 222 aa truncated C-terminal to the CARD. Some of these cDNA mutations are consistent with aberrant Ig somatic hypermutation (a physiological mechanism that results in affinity maturation of antibody after a B cell has encountered antigen, but that can also occur abnormally in other genes in B cell malignancies,63  whereas mutations that occur in polyadenine and polythymidine tracts of BCL10 are reminiscent of defective mismatch repair, as in replication error (RER)-positive malignancies.64  Microsatellite instability is rare in B cell NHL in general, but the replication error phenotype appears to be common in MALT lymphomas specifically.65  Studies of BCL10 CARD-truncation mutants revealed these proteins to be unable to induce death or activate NFκB, whereas C-terminal truncation mutants lost proapoptosis but retained NFκB activation.

These data suggested that BCL10 might normally function as a tumor suppressor (given its proapoptotic capabilities), whereas the truncation mutants thought to be overexpressed in t(1;14)-positive MALT tumors would be nonfunctional (CARD-truncation mutants) or might be capable of providing both antiapoptotic and pro-proliferative signals mediated by NFκB transcriptional targets such as cIAP-1 and -2, MYC and IL6 (C-terminal truncation mutants).66,67  However, the relative contributions of BCL10 mutants to MALT tumorigenesis are not clear and await further study. In addition, it appears that the effect of normal BCL10 on apoptosis may depend in part upon the developmental stage of a given cell and possibly the level of gene expression; for example, Yoneda et al have reported Bcl10 transgenic mice in which expression was driven by the ubiquitously active prion promoter, identifying enhanced apoptosis of T and B cells and atrophy of the thymus and spleen but no abnormal apoptosis in other organs such as brain in which Bcl10 was also aberrantly expressed at high levels.62  In marked contrast, transgenic mice in which BCL10 is linked to an immunoglobulin enhancer construct that directs expression to T and B cells only develop splenomegaly due to a dramatic and specific expansion of marginal zone B cells, creating a mouse model reminiscent of human splenic marginal zone lymphoma (S.W. Morris, manuscript submitted, and abstract at this meeting). Interestingly, mice generated using the same immunoglobulin enhancer-containing construct that express either a CARD-truncation or C-terminal truncation mutant (both cloned from a t(1;14)-positive MALT tumor) exhibit no evident abnormalities. These data bring into question the functional relevance of BCL10 truncation mutants in the genesis of MALT lymphoma, given that normal BCL10 gene expression under conditions that closely mimic the t(1;14) produces marginal zone B cell expansion.

The existence and functional importance of BCL10 mutations in other tumors is also unclear. Spurred by an initial report suggesting the presence of truncating BCL10 mutations independent of t(1;14) in a high percentage of MALT and other NHLs, and solid tumors such as mesotheliomas, and colon and testicular carcinomas,52  a large number of investigators have examined various malignancies for BCL10 mutation (representative publications include refs.68,69,70,71,72,73 ). Based on these manuscripts, BCL10 mutations in nonlymphoid hematopoietic malignancies and solid tumors of all types appear to be very rare and unlikely to play a significant role in oncogenesis. Further, in contrast to the high incidence of mutation initially reported in lymphoid malignancies (∼45% of B-and T-lymphomas), the combined data suggest that at most 5-10% of B cell NHLs may contain mutations, and mutation in T-lineage disease is rare. Also, monoallelic 1p22 deletions encompassing BCL10 have been detected by FISH and/or LOH analysis in a number of tumor types, but coding region mutations of the remaining allele are rare. Whether BCL10 haploinsufficiency might be sufficient to accelerate tumor progression, as shown for Bax, for example,74  is unknown. In summary, the initially reported high frequency of BCL10 mutation in a variety of hematopoietic and solid tumors has not been borne out by additional data from a large number of studies. Nonetheless, BCL10 mutation may contribute to the development of a small subset of lymphoid and possibly other malignancies. Additional study of the functional consequences of normal or mutant BCL10 expression using experimental in vivo models, and of the phenotype observed with Bcl10 absence, will be required to determine the exact role of this gene in oncogenesis.

Conclusion

Dramatic progress has been made in our understanding of non-Hodgkin lymphomagenesis through the characterization of recurrent chromosomal rearrangements. However, despite identification of many critical oncogenes, their functional consequences and exact mechanisms of lymphoid cell transformation remain to be fully elucidated. In the year 2000 and beyond, focus will shift toward these functional questions, as well as to the utilization of methods such as cDNA microarray analysis to more globally assess gene dysregulation in lymphomagenesis.75 

1
Pannetier C, Cochet M, Darche S, Casrouge A, Zoller M, Kourilsky P. The sizes of the CDR3 hypervariable regions of the murine T-cell receptor beta chains vary as a function of the recombined germ-line segments.
Proc Natl Acad Sci U S A
.
1993
;
90
:
4319
-4323.
2
Davi F, Faili A, Gritti C, et al. Early onset of immunoglobulin heavy chain gene rearrangements in normal human bone marrow CD34+ cells.
Blood
.
1997
;
90
:
4014
-4021.
3
Landman-Parker J, Aubin J, Delabesse E, et al. Simplified strategies for minimal residual disease detection in B-cell precursor acute lymphoblastic leukaemia.
Br J Haematol
.
1996
;
95
:
281
-290.
4
Deane M, Norton JD. Immunoglobulin gene `fingerprinting': an approach to analysis of B lymphoid clonality in lymphoproliferative disorders.
Br J Haematol
.
1991
;
77
:
274
-281.
5
Delabesse E, Burtin ML, Millien C, et al. Rapid, multifluorescent TCRG Vgamma and Jgamma typing: application to T cell acute lymphoblastic leukemia and to the detection of minor clonal populations [In Process Citation].
Leukemia
.
2000
;
14
:
1143
-1152.
6
Tierens A, Lozano MD, Wickert R, Chan WC, Greiner TC. High-resolution analysis of immunoglobulin heavy-chain gene rearrangements using denaturing gradient gel electrophoresis.
Diagn Mol Pathol
.
1996
;
5
:
159
-165.
7
Linke B, Pyttlich J, Tiemann M, et al. Identification and structural analysis of rearranged immunoglobulin heavy chain genes in lymphomas and leukemias.
Leukemia
.
1995
;
9
:
840
-847.
8
Baruchel A, Cayuela JM, Macintyre E, Berger R, Sigaux F. Assessment of clonal evolution at Ig/TCR loci in acute lymphoblastic leukaemia by single-strand conformation polymorphism studies and highly resolutive PCR derived methods: implication for a general strategy of minimal residual disease detection.
Br J Haematol
.
1995
;
90
:
85
-93.
9
Macintyre EA, d'Auriol L, Duparc N, Leverger G, Galibert F, Sigaux F. Use of oligonucleotide probes directed against T cell antigen receptor gamma delta variable-(diversity)-joining junctional sequences as a general method for detecting minimal residual disease in acute lymphoblastic leukemias.
J Clin Invest
.
1990
;
86
:
2125
-2135.
10
van Dongen JJ, Seriu T, Panzer-Grumayer ER, et al. Prognostic value of minimal residual disease in acute lymphoblastic leukaemia in childhood.
Lancet
.
1998
;
352
:
1731
-1738.
11
Cave H, van der Werff ten Bosch J, Suciu S, et al. Clinical significance of minimal residual disease in childhood acute lymphoblastic leukemia. European Organization for Research and Treatment of Cancer—Childhood Leukemia Cooperative Group.
N Engl J Med
.
1998
;
339
:
591
-598.
12
Pongers-Willemse MJ, Verhagen OJ, Tibbe GJ, et al. Real-time quantitative PCR for the detection of minimal residual disease in acute lymphoblastic leukemia using junctional region specific TaqMan probes.
Leukemia
.
1998
;
12
:
2006
-2014.
13
Donovan JW, Ladetto M, Zou G, et al. Immunoglobulin heavy-chain consensus probes for real-time PCR quantification of residual disease in acute lymphoblastic leukemia.
Blood
.
2000
;
95
:
2651
-2658.
14
Kuppers R, Klein U, Hansmann ML, Rajewsky K. Cellular origin of human B-cell lymphomas.
N Engl J Med
.
1999
;
341
:
1520
-1529.
15
Delfau MH, Hance AJ, Lecossier D, Vilmer E, Grandchamp B. Restricted diversity of V gamma 9-JP rearrangements in unstimulated human gamma/delta T lymphocytes.
Eur J Immunol
.
1992
;
22
:
2437
-2443.
16
Richards SJ, Short M, Scott CS. Clonal CD3+CD8+ large granular lymphocyte (LGL)/NK-associated (NKa) expansions: primary malignancies or secondary reactive phenomena?
Leuk Lymphoma
.
1995
;
17
:
303
-311.
17
Aisenberg AC. Utility of gene rearrangements in lymphoid malignancies.
Annu Rev Med
.
1993
;
44
:
75
-84.
18
Lorenzen J, Li G, Zhao-Hohn M, Wintzer C, Fischer R, Hansmann ML. Angioimmunoblastic lymphadenopathy type of T-cell lymphoma and angioimmunoblastic lymphadenopathy: a clinicopathological and molecular biological study of 13 Chinese patients using polymerase chain reaction and paraffin-embedded tissues.
Virchows Arch
.
1994
;
424
:
593
-600.
19
Hodges E, Quin CT, Wright DH, Smith JL. Oligoclonal populations of T and B cells in a case of angioimmunoblastic T-cell lymphoma predominantly infiltrated by T cells of the VB5.1 family.
Mol Pathol
.
1997
;
50
:
15
-17.
20
Stevenson F, Sahota S, Zhu D, et al. Insight into the origin and clonal history of B-cell tumors as revealed by analysis of immunoglobulin variable region genes.
Immunol Rev
.
1998
;
162
:
247
-259.
21
Aubin J, Davi F, Nguyen-Salomon F, et al. Description of a novel FR1 IgH PCR strategy and its comparison with three other strategies for the detection of clonality in B cell malignancies.
Leukemia
.
1995
;
9
:
471
-479.
22
Ng VL, Hurt MH, Herndier BG, Fry KE, McGrath MS. VH gene use by HIV type 1-associated lymphoproliferations.
AIDS Res Hum Retroviruses
.
1997
;
13
:
135
-149.
23
Derksen PW, Langerak AW, Kerkhof E, et al. Comparison of different polymerase chain reaction-based approaches for clonality assessment of immunoglobulin heavy-chain gene rearrangements in B-cell neoplasia.
Mod Pathol
.
1999
;
12
:
794
-805.
24
Bottaro M, Berti E, Biondi A, Migone N, Crosti L. Heteroduplex analysis of T-cell receptor gamma gene rearrangements for diagnosis and monitoring of cutaneous T-cell lymphomas.
Blood
.
1994
;
83
:
3271
-3278.
25
Cellier C, Patey N, Mauvieux L, et al. Abnormal intestinal intraepithelial lymphocytes in refractory sprue.
Gastroenterology
.
1998
;
114
:
471
-481.
26
Skinnider BF, Connors JM, Sutcliffe SB, Gascoyne RD. Anaplastic large cell lymphoma: a clinicopathologic analysis.
Hematol Oncol
.
1999
;
17
:
137
-148.
27
Kadin ME, Morris SW. The t(2;5) in human lymphomas.
Leuk Lymphoma
.
1998
;
29
:
249
-256.
28
Hilgenfeld E, Padilla-Nash H, Schrock E, Ried T. Analysis of B-cell neoplasias by spectral karyotyping (SKY).
Curr Top Microbiol Immunol
.
1999
;
246
:
169
-174.
29
Nordgren A, Sorensen AG, Tinggaard-Pedersen N, Blennow E, Larsson C, Lagercrantz S. New chromosomal breakpoints in non-Hodgkin's lymphomas revealed by spectral karyotyping and G-banding.
Int J Mol Med
.
2000
;
5
:
485
-492.
30
Golub TR, Barker GF, Bohlander SK, et al. Fusion of the TEL gene on 12p13 to the AML1 gene on 21q22 in acute lymphoblastic leukemia.
Proc Natl Acad Sci U S A
.
1995
;
92
:
4917
-4921.
31
Romana SP, Mauchauffe M, Le Coniat M, et al. The t(12;21) of acute lymphoblastic leukemia results in a tel-AML1 gene fusion.
Blood
.
1995
;
85
:
3662
-3670.
32
Richelda R, Ronchetti D, Baldini L, et al. A novel chromosomal translocation t(4;14)(p16.3;q32) in multiple myeloma involves the fibroblast growth-factor receptor 3 gene.
Blood
.
1997
;
90
:
4062
-4070.
33
Chesi M, Nardini E, Brents LA, et al. Frequent translocation t(4;14)(p16.3;q32.3) in multiple myeloma is associated with increased expression and activating mutations of fibroblast growth factor receptor 3.
Nat Genet
.
1997
;
16
:
260
-264.
34
Rack KA, Delabesse E, Radford-Weiss I, et al. Simultaneous detection of MYC, BVR1, and PVT1 translocations in lymphoid malignancies by fluorescence in situ hybridization.
Genes Chromosomes Cancer
.
1998
;
23
:
220
-226.
35
Wlodarska I, Mecucci C, Stul M, et al. Fluorescence in situ hybridization identifies new chromosomal changes involving 3q27 in non-Hodgkin's lymphomas with BCL6/LAZ3 rearrangement.
Genes Chromosomes Cancer
.
1995
;
14
:
1
-7.
36
Rack KA, Salomon-Nguyen F, Radford-Weiss I, et al. FISH detection of chromosome 14q32/IgH translocations: evaluation in follicular lymphoma.
Br J Haematol
.
1998
;
103
:
495
-504.
37
Vaandrager JW, Schuuring E, Raap T, Philippo K, Kleiverda K, Kluin P. Interphase FISH detection of BCL2 rearrangement in follicular lymphoma using breakpoint-flanking probes.
Genes Chromosomes Cancer
.
2000
;
27
:
85
-94.
38
Kluin PH, Schuuring E. FISH and related techniques in the diagnosis of lymphoma.
Cancer Surv
.
1997
;
30
:
3
-20.
39
Dalla-Favera R, Migliazza A, Chang CC, et al. Molecular Pathogenesis of B Cell Malignancy: the role of BCL6.
Curr Top Microbiol Immunol
.
1999
;
246
:
257
-263.
40
Ladanyi M, Wang S. Detection of rearrangements of the BCL2 major breakpoint region in follicular lymphomas. Correlation of polymerase chain reaction results with Southern blot analysis.
Diagn Mol Pathol
.
1992
;
1
:
31
-35.
41
Rimokh R, Berger F, Delsol G, et al. Detection of the chromosomal translocation t(11;14) by polymerase chain reaction in mantle cell lymphomas.
Blood
.
1994
;
83
:
1871
-1875.
42
Jadayel D, Matutes E, Dyer MJ, et al. Splenic lymphoma with villous lymphocytes: analysis of BCL-1 rearrangements and expression of the cyclin D1 gene.
Blood
.
1994
;
83
:
3664
-3671.
43
Avet-Loiseau H, Garand R, Gaillard F, et al. Detection of t(11;14) using interphase molecular cytogenetics in mantle cell lymphoma and atypical chronic lymphocytic leukemia.
Genes Chromosomes Cancer
.
1998
;
23
:
175
-182.
44
Uchimaru K, Taniguchi T, Yoshikawa M, et al. Detection of cyclin D1 (BCL1, PRAD1) overexpression by a simple competitive reverse transcription-polymerase chain reaction assay in t(11;14)(q13;q32)-bearing B-cell malignancies and/or mantle cell lymphoma.
Blood
.
1997
;
89
:
965
-974.
45
Mauvieux L, Canioni D, Hermine O, et al. Quantitative RNA slot-blot analysis of CCND1/cyclin D1 expression in suspected mantle cell lymphoma.
Leukemia
.
1998
;
12
:
78
-85.
46
Piris MA, Mollejo M, Campo E, Menarguez J, Flores T, Isaacson PG. A marginal zone pattern may be found in different varieties of non-Hodgkin's lymphoma: the morphology and immunohistology of splenic involvement by B-cell lymphomas simulating splenic marginal zone lymphoma.
Histopathology
.
1998
;
33
:
230
-239.
47
Troussard X, Mauvieux L, Radford-Weiss I, et al. Genetic analysis of splenic lymphoma with villous lymphocytes: a Groupe Francais d'Hematologie Cellulaire (GFHC) study.
Br J Haematol
.
1998
;
101
:
712
-721.
48
Ronchetti D, Finelli P, Richelda R, et al. Molecular analysis of 11q13 breakpoints in multiple myeloma.
Blood
.
1999
;
93
:
1330
-1337.
49
de Boer CJ, Vaandrager JW, van Krieken JH, Holmes Z, Kluin PM, Schuuring E. Visualization of mono-allelic chromosomal aberrations 3′ and 5′ of the cyclin D1 gene in mantle cell lymphoma using DNA fiber fluorescence in situ hybridization.
Oncogene
.
1997
;
15
:
1599
-1603.
50
Ladanyi M, Cavalchire G, Morris SW, Downing J, Filippa DA. Reverse transcriptase polymerase chain reaction for the Ki-1 anaplastic large cell lymphoma-associated t(2;5) translocation in Hodgkin's disease.
Am J Pathol
.
1994
;
145
:
1296
-1300.
51
Waggott W, Lo YM, Bastard C, et al. Detection of NPM-ALK DNA rearrangement in CD30 positive anaplastic large cell lymphoma.
Br J Haematol
.
1995
;
89
:
905
-907.
52
Lamant L, Meggetto F, al Saati T, et al. High incidence of the t(2;5)(p23;q35) translocation in anaplastic large cell lymphoma and its lack of detection in Hodgkin's disease. Comparison of cytogenetic analysis, reverse transcriptase-polymerase chain reaction, and P-80 immunostaining.
Blood
.
1996
;
87
:
284
-291.
53
Wlodarska I, De Wolf-Peeters C, Falini B, et al. The cryptic inv(2)(p23q35) defines a new molecular genetic subtype of ALK-positive anaplastic large-cell lymphoma.
Blood
.
1998
;
92
:
2688
-2695.
54
Mason DY, Pulford KA, Bischof D, et al. Nucleolar localization of the nucleophosmin-anaplastic lymphoma kinase is not required for malignant transformation.
Cancer Res
.
1998
;
58
:
1057
-1062.
55
Stec I, Wright TJ, van Ommen GJ, et al. WHSC1, a 90 kb SET domain-containing gene, expressed in early development and homologous to a Drosophila dysmorphy gene maps in the Wolf-Hirschhorn syndrome critical region and is fused to IgH in t(4;14) multiple myeloma [published erratum appears in Hum Mol Genet 1998 Sep;7(9):1527].
Hum Mol Genet
.
1998
;
7
:
1071
-1082.
56
Chesi M, Nardini E, Lim RS, Smith KD, Kuehl WM, Bergsagel PL. The t(4;14) translocation in myeloma dysregulates both FGFR3 and a novel gene, MMSET, resulting in IgH/MMSET hybrid transcripts.
Blood
.
1998
;
92
;
3025
-3034.
57
Janssen JW, Vaandrager JW, Heuser T, et al. Concurrent activation of a novel putative transforming gene, myeov, and cyclin D1 in a subset of multiple myeloma cell lines with t(11;14)(q13;q32).
Blood
.
2000
;
95
:
2691
-2698.
58
Alizadeh AA, Staudt LM. Genomic-scale gene expression profiling of normal and malignant immune cells.
Curr Opin Immunol
.
2000
;
12
:
219
-225.
59
Alizadeh AA, Eisen MB, Davis RE, et al. Distinct types of diffuse large B-cell lymphoma identified by gene expression profiling [see comments].
Nature
.
2000
;
403
:
503
-511.
60
Frank TS, Svoboda-Newman SM, Hsi ED. Comparison of methods for extracting DNA from formalin-fixed paraffin sections for nonisotopic PCR.
Diagn Mol Pathol
.
1996
;
5
:
220
-224.
61
Tbakhi A, Totos G, Pettay JD, Myles J, Tubbs RR. The effect of fixation on detection of B-cell clonality by polymerase chain reaction.
Mod Pathol
.
1999
;
12
:
272
-278.
62
Morgan GJ, Pratt G. Modern molecular diagnostics and the management of haematological malignancies.
Clin Lab Haematol
.
1998
;
20
:
135
-141.
63
Johnson PW, Swinbank K, MacLennan S, et al. Variability of polymerase chain reaction detection of the bcl-2-IgH translocation in an international multicentre study.
Ann Oncol
.
1999
;
10
:
1349
-1354.
64
Liu Y, Hernandez AM, Shibata D, Cortopassi GA. BCL2 translocation frequency rises with age in humans.
Proc Natl Acad Sci U S A
.
1994
;
91
:
8910
-8914.
65
Limpens J, Stad R, Vos C, et al. Lymphoma-associated translocation t(14;18) in blood B cells of normal individuals.
Blood
.
1995
;
85
:
2528
-2536.
66
van Dongen JJ, Macintyre EA, Gabert JA, et al. Standardized RT-PCR analysis of fusion gene transcripts from chromosome aberrations in acute leukemia for detection of minimal residual disease. Report of the BIOMED-1 Concerted Action: investigation of minimal residual disease in acute leukemia.
Leukemia
.
1999
;
13
:
1901
-1928.
67
Freedman AS, Neuberg D, Mauch P, et al. Long-term follow-up of autologous bone marrow transplantation in patients with relapsed follicular lymphoma.
Blood
.
1999
;
94
:
3325
-3333.
68
Gribben JG. Stem-cell transplantation for indolent lymphoma.
Semin Hematol
.
1999
;
36
:
18
-25.
69
Corradini P, Ladetto M, Pileri A, Tarella C. Clinical relevance of minimal residual disease monitoring in non-Hodgkin's lymphomas: a critical reappraisal of molecular strategies.
Leukemia
.
1999
;
13
:
1691
-1695.
70
Corradini P, Astolfi M, Cherasco C, et al. Molecular monitoring of minimal residual disease in follicular and mantle cell non-Hodgkin's lymphomas treated with high-dose chemotherapy and peripheral blood progenitor cell autografting.
Blood
.
1997
;
89
:
724
-731.
71
Aiello A, Giardini R, Tondini C, et al. PCR-based clonality analysis: a reliable method for the diagnosis and follow-up monitoring of conservatively treated gastric B-cell MALT lymphomas?
Histopathology
.
1999
;
34
:
326
-330.
1
Tycko B, Sklar J. Chromosomal translocations in lymphoid neoplasia: A reappraisal of the recombinase model.
Cancer Cells
.
1990
;
2
:
1
-8.
2
Rabbitts TH. Chromosomal translocations in human cancer.
Nature
.
1994
;
372
:
143
-149.
3
Kuppers R, Klein U, Hansmann ML, Rajewsky K. Cellular origin of human B-cell lymphomas.
N Engl J Med
.
1999
;
341
:
1520
-1529.
4
Vanasse G, Concannon P, Willerford DM. Regulated genomic instability and neoplasia in the lymphoid lineage.
Blood
.
1999
;
94
:
3997
-4010.
5
Alt FW, Oltz EM, Young F, Gorman J, Taccioli G, Chen J. VDJ recombination.
Immunology Today
.
1992
;
13
:
313
.
6
Taccioli GE, Alt FW. Potential targets for autosomal SCID mutations.
Curr Opin Immunol
.
1992
;
7
:
436
-440.
7
McBlane JF, van Gent DC, Ramsden DA, Romeo C, Cuomo CA, Oettinger MA. Cleavage at a V(D)J recombination signal requires only RAG1 and RAG2 proteins and occurs in two steps.
Cell
.
1995
;
83
:
387
-395.
8
Hiom K, Gellert M. Assembly of a 12/23 paired signal complex: a critical control point in V(D)J recombination.
Mol Cell
.
1998
;
1
:
1011
-1019.
9
Schwartz K, Gauss GH, Ludwig L, et al. RAG mutations in human B cell-negative SCID.
Science
.
1996
;
274
:
97
-99.
10
Sleckman BP, Gorman J, Alt FW. Accessibility control of antigen-receptor variable-region gene assembly: role of cis-acting elements.
Annu Rev Immunol
.
1996
;
14
:
459
-481.
11
Jager U, Bocskor S, Le T, et al. Follicular lymmphomas' BCL-2/IgH junctions contain templated nucleotide insertions: novel insights into the machanism of t(14;18) translocations.
Blood
.
2000
;
95
:
3520
-9.
12
Agrawal A, Eastman QM, Schatz DG. Transposition mediated by RAG1 and RAG2 and its implications for the evolution of the immune system.
Nature
.
1998
;
394
:
744
-751.
13
Hiom K, Melek M, Gellert M. DNA transposition by the RAG1 and RAG2 proteins: a possible source of oncogenic translocations.
Cell
.
1998
;
94
:
463
-470.
14
Stavnezer J. Antibody class switching.
Adv Immunol
.
1996
;
61
:
79
-146.
15
Snapper CM, Marcu KB, Zelazowski P. The immunoglobulin class switch: beyond “accessibility”.
Immunity
.
1997
;
6
:
217
-223.
16
Tracy RB, Hsieh CL, Lieber MR. Stable RNA/DNA hybrids in the mammalian genome: inducible intermediates in immunoglobulin class switch recombination.
Science
.
2000
;
288
:
1058
-1061.
17
Iwasato T, Shimizu A, Honjo T, Yamagishi H. Circular DNA is excised by immunoglobulin class switch recombination.
Cell
.
1990
;
62
:
143
-149.
18
Rolink A, Melchers F, Andersson J. The SCID but not the RAG2 gene product is required for S mu-S epsilon heavy chain class switching.
Immunity
.
1996
;
5
:
319
-330.
19
Manis JP, Gu Y, Lansford R, et al. Ku70 is required for late B cell development and immunoglobulin heavy chain class switching.
J Exp Med
.
1998
;
187
:
2081
-2089.
20
Casellas R, Nussenzweig A, Wuerffel R, et al. Ku80 is required for immunoglobulin isotype switching.
Embo J
.
1998
;
17
:
2404
-2411.
21
Neri A, Barriga F, Knowles DM, Magrath IT, Dalla-Favera R. Different regions of the immunoglobulin heavy-chain locus are involved in chromosomal translocations in distinct pathogenetic forms of Burkitt lymphoma.
Proc Nat Acad Sci USA
.
1988
;
85
:
2748
-2752.
22
Shiramizu B, Barriga F, Neequaye J, et al. Patterns of chromosomal breakpoint locations in Burkitt's lymphoma: relevance to geography and Epstein-Barr virus association.
Blood
.
1991
;
77
:
1516
-1526.
23
Potter M. Experimental plasmacytomagenesis in mice.
Hematol Oncol Clin North Am
.
1997
;
11
:
323
-347.
24
Gabrea A, Bergsagel PL, Chesi M, Shou Y, Kuehl WM. Insertion of Excised IgH Switch Sequences Causes Overexpression of Cyclin D1 in a Myeloma Tumor Cell.
Molecular Cell
.
1999
;
3
:
119
-123.
25
Storb U. Progress in understanding the mechanism and consequences of somatic hypermutation.
Immunol Rev
.
1998
;
162
:
5
-11.
26
Kelsoe G. V(D)J hypermutation and DNA mismatch repair: vexed by fixation.
Proc Nat Acad Sci USA
.
1998
;
95
:
6576
-6577.
27
Goossens T, Klein U, Kuppers R. Frequent occurrence of deletions and duplications during somatic hypermutation: implications for oncogene translocations and heavy chain disease.
Proc Natl Acad Sci USA
.
1998
;
95
:
2463
-2468.
28
Wilson PC, de Bouteiller O, Liu YJ, et al. Somatic hypermutation introduces insertions and deletions into immunoglobulin V genes.
J Exp Med
.
1998
;
187
:
59
-70.
29
Sale JE, Neuberger MS. TdT-accessible breaks are scattered over the immunoglobulin V domain inn a constitutively hypermutating B cell line.
Immunity
.
1998
;
9
:
859
-869.
30
Klein U, Goossens T, Fischer M, et al. Somatic hypermutation in normal and transformed B cells.
Immunol Rev
.
1998
;
162
:
261
-280.
31
Rabbitts TH, Forster A, Hamlyn P, Baer R. Effect of somatic mutation within translocated c-myc genes in Burkitt's lymphoma.
Nature
.
1984
;
309
:
592
-597.
32
Migliazza A, Martinotti S, Chen W, et al. Frequent somatic hypermutation of the 5′ noncoding region of the BCL6 gene in B-cell lymphoma.
Proc Natl Acad Sci U S A
.
1995
;
92
:
12520
-12524.
33
Shen HM, Peters A, Baron B, Zhu X, Storb U. Mutation of BCL-6 gene in normal B cells by the process of somatic hypermutation of Ig genes.
Science
.
1998
;
280
:
1750
-1752.
34
Pasqualucci L, Migliazza A, Fracchiolla N, et al. BCL-6 mutations in normal germinal center B cells: evidence of somatic hypermutation acting outside Ig loci.
Proc Natl Acad Sci U S A
.
1998
;
95
:
11816
-11821.
35
Jacob J, Kelsoe G, Rajewsky K, Weiss U. Intraclonal generation of antibody mutants in germinal centres.
Nature
.
1991
;
354
:
389
-392.
36
Liu YJ, Arpin C. Germinal center development.
Immunol Rev
.
1997
;
156
:
111
-126.
37
Liu YJ, Malisan F, de Bouteiller O, et al. Within germinal centers, isotype switching of immunoglobulin genes occurs after the onset of somatic mutation.
Immunity
.
1996
;
4
:
241
-250.
38
Nussenzweig MC. Immune receptor editing: revise and select.
Cell
.
1998
;
95
:
875
-878.
39
Chen C, Prak EL, Weigert M. Editing disease-associated autoantibodies.
Immunity
.
1997
;
6
:
97
-105.
40
Han S, Dillon SR, Zheng B, Shimoda M, Schlissel MS, Kelsoe G. V(D)J recombinase activity in a subset of germinal center B lymphocytes.
Science
.
1997
;
278
:
301
-305.
41
Papavasiliou F, Casellas R, Suh H, et al. V(D)J recombination in mature B cells: a mechanism for altering antibody responses.
Science
.
1997
;
278
:
298
-301.
42
Yu W, Nagaoka H, Jankovic M, et al. Continued RAG expression in late stages of B cell development and no apparent reinduction after immunization.
Nature
.
1999
;
400
:
682
-687.
43
Monroe RJ, Seidl KJ, Gaertner F, et al. Rag2:GFP knockin mice reveal novel aspects of Rag2 expression in primary and peripheral lymphoid tissues.
Immunity
.
1999
;
11
:
201
-212.
44
McMahan CJ, Fink PJ. RAG reexpression and DNA recombination at T cell receptor loci in peripheral CD4+ T cells.
Immunity
.
1998
;
9
:
637
-647.
45
Harris NL, Jaffe ES, Stein H, et al. A revised European-American classification of lymphoid neoplasms: a proposal from the International Lymphoma Study Group.
Blood
.
1994
;
84
:
1361
-1392.
46
Isaacson PG. Gastric MALT lymphoma: from concept to cure.
Ann Oncol
.
1999
;
10
:
637
-645.
47
Ng VL, Hurt MH, Fein CL, et al. IgMs produced by two acquired immune deficiency syndrome lymphoma cell lines: Ig binding specificity and VH-gene putative somatic mutation analysis.
Blood
.
1994
;
83
:
1067
-1078.
48
Cunto-Amesty G, Przybylski G, Honczarenko M, Monroe JG, Silberstein LE. Evidence that immunoglobulin specificities of AIDS-related lymphoma are not directed to HIV-related antigens.
Blood
.
2000
;
95
:
1393
-1399.
49
Gatti RA, Ataxia-telangiectasia. In: eds.
The metabolic and molecular bases for inherited disease
. Vol New York, N.Y.: McGraw-Hill;
1998
:
50
Savitsky K, Bar-Shira A, Gilad S, et al. A single ataxia telangiectasia gene with a product similar to PI-3 kinase.
Science
.
1995
;
268
:
1749
-1753.
51
Kojis TL, Gatti RA, Sparkes RS. The cytogenetics of ataxia telangiectasia.
Cancer Genet Cytogenet
.
1991
;
56
:
143
-156.
52
Taylor AM, Metcalfe JA, Thick J, Mak YF. Leukemia and lymphoma in ataxia telangiectasia.
Blood
.
1996
;
87
:
423
-438.
53
Morrell D, Cromartie E, Swift M. Mortality and cancer incidence in 263 patients with ataxia-telangiectasia.
J Natl Cancer Inst
.
1986
;
77
:
89
-92.
54
Virgilio L, Narducci MG, Isobe M, et al. Identification of the TCL1 gene involved in T-cell malignancies.
Proc Natl Acad Sci U S A
.
1994
;
91
:
12530
-12534.
55
Wegner RD, Chrzanowska KH, Spreling K, Stumm M. Ataxiatelangiectasia variants (Nijmegen Breakage Syndrome). In:
Primary Immunodeficiency Diseases, A Molecular and Genetic Approach
. Vol Oxford: Oxford University Press;
1999
:
56
Carney JP, Mase RS, Olivares H, et al. The hMre 11/hRad50 protein complex and Nijmegen Breakage Syndrome: linkage of double-strand break repair to the cellular DNA damage response.
Cell
.
1998
;
93
:
477
-486.
57
Varon R, Vissinga C, Platzer M, et al. Nibrin, a novel DNA double-strand break repair protein is mutated in Nijmegen Breakage Syndrome.
Cell
.
1998
;
93
:
467
-476.
58
Ellis NA, Groden J, Ye TZ, et al. The Bloom's syndrome gene product is homologous to RecQ helicases.
Cell
.
1995
;
83
:
655
-666.
59
Ganesh A, North P, Thacker J. Repair and misrepair of site-specific DNA double-strand breaks by human cell extracts.
Mutat Res
.
1993
;
299
:
251
-259.
60
Hsieh CL, Arlett CF, Lieber MR. V(D)J recombination in ataxia telangiectasia, Bloom's syndrome, and a DNA ligase I-associated immunodeficiency disorder.
J Biol Chem
.
1993
;
268
:
20105
-20109.
61
Matsuoka S, Huang M, Elledge SJ. Linkage of ATM to cell cycle regulation by the Chk2 protein kinase.
Science
.
1998
;
282
:
1893
-1897.
62
Canman CE, Lim DS, Cimprich KA, et al. Activation of the ATM kinase by ionizing radiation and phosphorylation of p53.
Science
.
1998
;
281
:
1677
-1679.
63
Banin S, Moyal L, Shieh S, et al. Enhanced phosphorylation of p53 by ATM in response to DNA damage.
Science
.
1998
;
281
:
1674
-1677.
64
Gatei M, Young D, Cerosaletti KM, et al. ATM-dependent phosphorylation of nibrin in response to radiation exposure.
Nat Genet
.
2000
;
25
:
115
-119.
65
Lim DS, Kim ST, Xu B, et al. ATM phosphorylates p95/nbsl in an S-phase checkpoint pathway.
Nature
.
2000
;
404
:
613
-617.
66
Donehower LA, Harvey M, Slagle BL, et al. Mice deficient for p53 are developmentally normal but susceptible to spontaneous tumours.
Nature
.
1992
;
356
:
215
-221.
67
Bogue MA, Zhu C, Aguilar-Cordova E, Donehower LA, Roth DB. p53 is required for both radiation-induced differentiation and rescue of V(D)J rearrangement in scid mouse thymocytes.
Genes Dev
.
1996
;
10
:
553
-565.
68
Guidos CJ, Williams CJ, Grandal I, Knowles G, Huang MTF, Danska JS. V(D)J recombination activates a p53-dependent DNA damage checkpoint in scid lymphocyte precursors.
Genes Dev
.
1996
;
10
:
2038
-2054.
69
Nacht M, Strasser A, Chan YR, et al. Mutations in the p53 and SCID genes cooperate in tumorigenesis.
Genes Dev
.
1996
;
10
:
2055
-2066.
70
Vanasse GJ, Halbrook J, Thomas S, et al. Genetic pathway to recurrent chromosome translocations in murine lymphoma involves V(D)J recombinase.
J Clin Invest
.
1999
;
103
:
1669
-1675.
71
Lim DS, Vogel H, Willerford DM, Sands AT, Platt KA, Hasty P. Analysis of ku80-mutant mice and cells with deficient levels of p53.
Mol Cell Biol
.
2000
;
20
:
3772
-3780.
72
Gao Y, Ferguson DO, Xie W, et al. Interplay of p53 and DNA-repair protein XRCC4 in tumorigenesis, genomic stability and development.
Nature
.
2000
;
404
:
897
-900.
73
Difilippantonio MJ, Zhu J, Chen HT, et al. DNA repair protein Ku80 suppresses chromosomal aberrations and malignant transformation.
Nature
.
2000
;
404
:
510
-514.
1
Weisenburger DD, Armitage JO. Mantle cell lymphoma—an entity comes of age.
Blood
.
1996
;
87
:
4483
-4494.
2
Yang E, Korsmeyer SJ. Molecular thanatopsis: a discourse on the BCL2 family and cell death.
Blood
.
1996
;
88
:
386
-401.
3
Nesbit CE, Tersak JM, Prochownik EV. MYC oncogenes and human neoplastic disease.
Oncogene
.
1999
;
18
:
3004
-3016.
4
Luque I, Gelinas C. Rel/NF-kappa B and I kappa B factors in oncogenesis.
Semin Cancer Biol
.
1997
;
8
:
103
-111.
5
Dalla-Favera R, Migliazza A, Chang CC, et al. Molecular pathogenesis of B cell malignancy: the role of BCL-6.
Curr Top Microbiol Immunol
.
1999
;
246
:
257
-63.
6
Chan JK. Anaplastic large cell lymphoma: redefining its morphologic spectrum and importance of recognition of the ALK-positive subset [published erratum appears in Adv Anat Pathol 1998 Nov;5(6):399].
Adv Anat Pathol
.
1998
;
5
:
281
-313.
7
Kinney MC, Kadin ME. The pathologic and clinical spectrum of anaplastic large cell lymphoma and correlation with ALK gene dysregulation.
Am J Clin Pathol
.
1999
;
111
:
S56
-S67.
8
Mansouri A, Goudreau G, Gruss P. Pax genes and their role in organogenesis.
Cancer Res
.
1999
;
59
:
1707s
-1709s.
9
Schafer BW. Emerging roles for PAX transcription factors in cancer biology.
Gen Physiol Biophys
.
1998
;
17
:
211
-224.
10
Iida S, Rao PH, Ueda R, Chaganti RS, Dalla-Favera R. Chromosomal rearrangement of the PAX-5 locus in lymphoplasmacytic lymphoma with t(9;14)(p13;q32).
Leuk Lymphoma
.
1999
;
34
:
25
-33.
11
Hagman J, Wheat W, Fitzsimmons D, Hodsdon W, Negri J, Dizon F. Pax-5/BSAP: regulator of specific gene expression and differentiation in B lymphocytes.
Curr Top Microbiol Immunol
.
2000
;
245
:
169
-194.
12
Urbanek P, Wang ZQ, Fetka I, Wagner EF, Busslinger M. Complete block of early B cell differentiation and altered patterning of the posterior midbrain in mice lacking Pax5/BSAP.
Cell
.
1994
;
79
:
901
-912.
13
Wakatsuki Y, Neurath MF, Max EE, Strober W. The B cells-pecific transcription factor BSAP regulates B cell proliferation.
J Exp.Med
.
1994
;
179
:
1099
-1108.
14
Morrison AM, Nutt SL, Thevenin C, Rolink A, Busslinger M. Loss- and gain-of-function mutations reveal an important role of BSAP (Pax-5) at the start and end of B cell differentiation.
Semin.Immunol
.
1998
;
10
:
133
-142.
15
Libermann TA, Pan Z, Akbarali Y, et al. AML1 (CBFalpha2) cooperates with B cell-specific activating protein (BSAP/PAX5) in acitvation of the B cell-specific BLK gene promoter.
J Biol Chem
.
1999
;
274
:
24671
-24676.
16
Malek SN, Dordai DI, Reim J, Dintzis H, Desiderio S. Malignant transformation of early lymphoid progenitors in mice expressing an activated Blk tyrosine kinase.
Proc Natl Acad Sci U S A
.
1998
;
95
:
7351
-7356.
17
Texido G, Su IH, Mecklenbrauker I, et al. The B-cell-specific Src-family kinase Blk is dispensable for B-cell development and activation.
Mol Cell Biol
.
2000
;
20
:
1227
-1233.
18
Morrison AM, Jager U, Chott A, Schebesta M, Haas OA, Busslinger M. Deregulated PAX-5 transcription from a translocated IgH promoter in marginal zone lymphoma.
Blood
.
1998
;
92
:
3865
-3878.
19
Barr FG. The role of chimeric paired box transcription factors in the pathogenesis of pediatric rhabdomyosarcoma.
Cancer Res
.
1999
;
59
:
1712
-1715s.
20
Dyomin VG, Rao PH, Dalla-Favera R, Chaganti RK. BCL8, a novel gene involved in translocations affecting band 15q11-13 in diffuse large-cell lymphoma.
Proc Natl Acad Sci U S A
.
1997
;
94
:
5728
-5732.
21
Mitelman F, Mertens F, Johansson B. A breakpoint map of recurrent chromosomal rearrangements in human neoplasia.
Nat Genet
.
1997
;
15
:
417
-474.
22
Offit K, Wong G, Filippa DA, Tao Y, Chaganti RS. Cytogenetic analysis of 434 consecutively ascertained specimens of non-Hodgkin's lymphoma: clinical correlations.
Blood
.
1991
;
77
:
1508
-1515.
23
Willis TG, Zalcberg IR, Coignet LJA, et al.: Molecular cloning of translocation t(1;14)(q21;q32) defines a novel gene (BCL9) at chromosome 1q21.
Blood
.
1998
;
91
:
1873
-1881.
24
Dyomin VG, Palanisamy N, Lloyd KO, et al. MUC1 is activated in a B-cell lymphoma by the t(1;14)(q21;q23) translocation and is rearranged and amplified in B-cell lymphoma subsets.
Blood
.
2000
;
95
:
2666
-2671.
25
Gilles F, Goy A, Remache Y, Shue P, Zelenetz AD. MUC1 dysregulation as the consequence of a t(1;14)(q21;q32) translocation in an extranodal lymphoma.
Blood
.
2000
;
95
:
2930
-2936.
26
Spicer AP, Rowse GJ, Lidner TK, Gendler SJ. Delayed mammary tumor progression in Muc-1 null mice.
J Biol Chem
.
1995
;
270
:
30093
-30101.
27
Wesseling J, van der Valk SW, Vos HL, Sonnenberg A, Hilkens J. Episialin (MUC1) overexpression inhibits integrin-mediated cell adhesion to extracellular matrix components.
J Cell Biol
.
1995
;
129
:
255
-265.
28
Callanan MB, Le Baccon P, Mossuz P, et al. The IgG Fc receptor, FcgammaRIIB, is a target for deregulation by chromosomal translocation in malignant lymphoma.
Proc Natl Acad Sci USA
.
2000
;
97
:
309
-314.
29
Daeron M. Fc receptor biology.
Annu Rev Immunol
.
1997
;
15
:
203
-34.
30
de Andres B, Mueller AL, Verbeek S, Sandor M, Lynch RG. A regulatory role for Fcgamma receptors CD16 and CD32 in the development of murine B cells.
Blood
.
1998
;
92
:
2823
-2829.
31
Zusman T, Lisansky E, Arons E, et al. Contribution of the intracellular domain of murine Fc-gamma receptor type IIB1 to its tumor-enhancing potential.
Int J Cancer
.
1996
;
68
:
219
-227.
32
Hatzivassiliou G, Iida S, Rao PH, et al. Two novel genes related to the Fc receptor family involved in the t(1;14)(q21;32) chromosomal translocation in multiple myeloma.
Blood
.
1998
;
92
:
509a
. (abstr)
33
Harris NL, Jaffe ES, Stein H, et al. A revised European-American classification of lymphoid neoplasms: a proposal from the International Lymphoma Study Group.
Blood
.
1994
;
84
:
1361
-1392.
34
Campo E, Miquel R, Krenacs L, Sorbara L, Raffeld M, Jaffe ES. Primary nodal marginal zone lymphomas of splenic and MALT type.
Am J Surg Pathol
.
1999
;
23
:
59
-68.
35
Nathwani BN, Anderson JR, Armitage JO, et al. Marginal zone B-cell lymphoma: a clinical comparison of nodal and mucosa-associated lymphoid tissue types. Non-Hodgkin's Lymphoma Classification Project.
J Clin Oncol
.
1999
;
17
:
2486
-2492.
36
Isaacson PG. Mucosa-associated lymphoid tissue lymphoma.
Semin Hematol
.
1999
;
36
:
139
-147.
37
Wotherspoon AC. Gastric lymphoma of mucosa-associated lymphoid tissue and Helicobacter pylori.
Annu Rev Med
.
1998
;
49
:
289
-99:289-299.
38
Dierlamm J, Baens M, Wlodarska I, et al. The apoptosis inhibitor gene API2 and a novel 18q gene, MLT, are recurrently rearranged in the t(11;18)(q21;q21) associated with mucosa-associated lymphoid tissue lymphomas.
Blood
.
1999
;
93
:
3601
-3609.
39
Akagi T, Motegi M, Tamura A, et al. A novel gene, MALT1 at 18q21, is involved in t(11;18) (q21;q21) found in low-grade B-cell lymphoma of mucosa-associated lymphoid tissue.
Oncogene
.
1999
;
18
:
5785
-5794.
40
Morgan JA, Yin Y, Borowsky AD, et al. Breakpoints of the t(11;18)(q21;q21) in mucosa-associated lymphoid tissue (MALT) lymphoma lie within or near the previously undescribed gene MALT1 in chromosome 18.
Cancer Res
.
1999
;
59
:
6205
-6213.
41
Deveraux QL, Reed JC. IAP family proteins—suppressors of apoptosis.
Genes Dev
.
1999
;
13
:
239
-252.
42
Hofmann K, Bucher P, Tschopp J. The CARD domain: a new apoptotic signaling motif.
Trends Biochem Sci
.
1997
;
22
:
155
-156.
43
Saurin AJ, Borden KL, Boddy MN, Freemont PS. Does this have a familiar RING?
Trends Biochem Sci
.
1996
;
21
:
208
-214.
44
Roy N, Deveraux Q, Takahashi R, Salvesen G, Reed J. The c-IAP-1 and c-IAP-2 proteins are direct inhibitors of specific caspases.
EMBO J
.
1997
;
16
,
6914
-6925.
45
Hay BA, Wassarman DA, Rubin GM. Drosophila homologs of baculovirus inhibitor of apoptosis proteins function to block cell death.
Cell
.
1995
;
83
:
1253
-1262.
46
Knorr C, Amrehn C, Seeberger H, et al. Expression of costimulatory molecules in low-grade mucosa-associated lymphoid tissue-type lymphomas in vivo.
Am J Pathol
.
1999
;
155
:
2019
-2027.
47
Motegi M, Yonezumi M, Suzuki H, et al. API2-MALT1 chimeric transcripts involved in mucosa-associated lymphoid tissue type lymphoma predict heterogeneous products.
Am J Pathol
.
2000
;
156
:
807
-812.
48
Baens M, Maes B, Steyls A, Geboes K, Marynen P, De Wolf-Peeters C. The product of the t(11;18), an API2-MLT fusion, marks nearly half of gastric MALT type lymphomas without large cell proliferation.
Am J Pathol
.
2000
;
156
:
1433
-1439.
49
Remstein ED, James CD, Kurtin PJ. Incidence and subtype specificity of AP12-MALT1 fusion translocations in extranodal, nodal, and splenic marginal zone lymphomas.
Am J Pathol
.
2000
;
156
:
1183
-1188.
50
Rosenwald A, Ott G, Stilgenbauer S, et al. Exclusive detection of the t(11;18)(q21;q21) in extranodal marginal zone B cell lymphomas (MZBL) of MALT type in contrast to other MZBL and extranodal large B cell lymphomas.
Am J Pathol
.
1999
;
155
:
1817
-1821.
51
Wotherspoon AC, Pan L, Diss TC, Isaacson PG. Cytogenetic study of B-cell lymphoma of mucosa-associated lymphoid tissue.
Cancer Genet. Cytogenet
.
1991
;
58
:
35
-38.
52
Willis TG, Jadayel DM, Du MQ, et al. Bcl10 is involved in t(1;14)(p22;q32) of MALT B cell lymphoma and mutated in multiple tumor types.
Cell
.
1999
;
96
:
35
-45.
53
Zhang Q, Siebert R, Yan M, et al. Inactivating mutations and overexpression of BCL10, a caspase recruitment domain-containing gene, in MALT lymphoma with t(1;14)(p22;q32).
Nat Genet
.
1999
;
22
:
63
-68.
54
Li P, Nijhawan D, Budihardjo I, et al. Cytochrome c and dATP-dependent formation of Apaf-1/caspase-9 complex initiates an apoptotic protease cascade.
Cell
.
1997
;
91
:
479
-489.
55
Yang X, Chang HY, Baltimore D. Essential role of CED-4 oligomerization in CED-3 activation and apoptosis.
Science
.
1998
;
281
:
1355
-1357.
56
Koseki T, Inohara N, Chen S, et al. CIPER, a novel NF kappaB-activating protein containing a caspase recruitment domain with homology to herpesvirus-2 protein E10.
J Biol Chem
.
1999
;
274
:
9955
-9961.
57
Srinivasula SM, Ahmad M, Lin JH, et al. CLAP, a novel caspase recruitment domain-containing protein in the tumor necrosis factor receptor pathway, regulates NF-kappaB activation and apoptosis.
J Biol Chem
.
1999
;
274
:
17946
-17954.
58
Thome M, Martinon F, Hofmann K, et al. Equine herpesvirus-2 E10 gene product, but not its cellular homologue, activates NF-kappaB transcription factor and c-Jun N-terminal kinase.
J Biol Chem
.
1999
;
274
:
9962
-9968.
59
Yan M, Lee J, Schilbach S, Goddard A, Dixit V. mE10, a novel caspase recruitment domain-containing proapoptotic molecule.
J Biol Chem
.
1999
;
274
:
10287
-10292.
60
Telford EAR, Watson MS, Aird HC, Perry J, Davison AJ. The DNA sequence of equine herpesvirus 2.
J Mol Biol
.
1995
;
249
:
520
-528.
61
Costanzo A, Guiet C, Vito P. c-E10 is a caspase-recruiting domain-containing protein that interacts with components of death receptors signaling pathway and activates nuclear factor-kappaB.
J Biol Chem
.
1999
;
274
:
20127
-20132.
62
Yoneda T, Imaizumi K, Maeda M, et al. Regulatory mechanisms of TRAF2-mediated signal transduction by Bc110, a MALT lymphoma-associated protein.
J Biol Chem
.
2000
;
275
:
11114
-11120.
63
Shen HM, Peters A, Baron B, Zhu X, Storb U. Mutation of BCL-6 gene in normal B cells by the process of somatic hypermutation of Ig genes.
Science
.
1998
;
280
:
1750
-1752.
64
Rampino N, Yamamoto H, Ionov Y, et al. Somatic frameshift mutations in the BAX gene in colon cancers of the microsatellite mutator phenotype.
Science
.
1997
;
275
:
967
-969, 1997
65
Peng H, Chen G, Du M, Singh N, Isaacson PG, Pan L. Replication error phenotype and p53 gene mutation in lymphomas of mucosa-associated lymphoid tissue.
Am. J. Clin. Pathol
.
1996
;
148
:
643
-648.
66
Wang CY, Mayo MW, Korneluk RG, Goeddel DV, Baldwin ASJ. NF-kappaB antiapoptosis: induction of TRAF1 and TRAF2 and c-IAP1 and c-IAP2 to suppress caspase-8 activation.
Science
.
1998
;
281
:
1680
-1683.
67
Sonenshein GE. Rel/NF-κB transcription factors and the control of apoptosis.
Semin Cancer Biol
.
1997
;
8
:
113
-119.
68
Gill S, Broni J, Jefferies S, et al. BCL10 is rarely mutated in human prostate carcinoma, small-cell lung cancer, head and neck tumours, renal carcinoma and sarcomas.
Br J Cancer
.
1999
;
80
:
1565
-1568.
69
Lee SH, Shin MS, Kim HS, et al. Point mutations and deletions of the Bc110 gene in solid tumors and malignant lymphomas.
Cancer Res
.
1999
;
59
:
5674
-5677.
70
Luminari S, Intini D, Baldini L, et al. BCL10 gene mutations rarely occur in lymphoid malignancies.
Leukemia
.
2000
;
14
:
905
-908.
71
Simms LA, Young J, Wicking C, Meltzer SJ, Jass JR, Leggett BA. The apoptotic regulatory gene, BCL10, is mutated in sporadic mismatch repair deficient colorectal cancers.
Cell Death Differ
.
2000
;
7
:
236
-237.
72
Du MQ, Peng H, Liu H, et al. BCL10 gene mutation in lymphoma.
Blood
.
2000
;
95
:
3885
-3890.
73
Takahashi H, Maeda Y, Seto M, Hosokawa Y. Nucleotide insertions and deletions within the homopolymeric runs of adenines and thymidines of BCL10 cDNAs in normal peripheral blood leukocytes.
Blood
.
2000
;
95
:
2728
-2729.
74
Shibata MA, Liu ML, Knudson MC, et al. Haploid loss of bax leads to accelerated mammary tumor development in C3(1)/SV40-TAg transgenic mice: reduction in protective apoptotic response at the preneoplastic stage.
EMBO J
.
1999
;
18
:
2692
-2701.
75
Alizadeh AA, Eisen MB, Davis RE, et al. Distinct types of diffuse large B-cell lymphoma identified by gene expression profiling.
Nature
.
2000
;
403
:
503
-511.