Skip to main content
Top

16-02-2016 | Hematologic cancers | Article

Risk Stratification in Multiple Myeloma

Journal: Current Hematologic Malignancy Reports

Authors: Melissa Gaik-Ming Ooi, Sanjay de Mel, Wee Joo Chng

Publisher: Springer US

Abstract

There are many prognostic variables in multiple myeloma and the difficulty is in deciding which is truly significant. The widely used International Staging System (ISS) does not incorporate genetics, age, and other important variables in its risk stratification. Although it has its own limitations, the recently published Revised International Staging System (R-ISS) that was built upon the framework of ISS, is a more comprehensive and predictive tool for multiple myeloma patients and should be henceforth utilised. We will review the current prognostic variables and their significance in this paper.

Introduction

Multiple Myeloma (MM) is a clonal plasma cell disorder characterised by the secretion of a monoclonal protein in the majority of patients. MM shows a variable clinical presentation ranging from the premalignant condition of monoclonal gammopathy of undetermined significance (MGUS) to smouldering MM, and symptomatic MM, extra-medullary myeloma or plasma cell leukaemia. The life expectancy of patients with this disease has improved with the advent of novel therapies. However, MM remains a highly heterogeneous disease with the overall survival (OS) ranging from a few months to decades among different cases. This variability results from the heterogeneity in the tumour and bone marrow microenvironment.
Prognostic factors in MM can be divided into tumour factors or host factors. Tumour biology factors relate to the MM chromosomal abnormalities, gene expression profile, and tumour burden. Lactate dehydrogenase (LDH), plasma cell proliferative rate, extramedullary disease and initial presentation as plasma cell leukaemia are additional measures of tumour burden. Host related factors include age, performance status, and comorbidities. Response to therapy is a factor related to both host and tumour biology.

Tumour Factors

Tumour Biology

Cytogenetic Abnormalities

Chromosomal changes are present in the plasma cells of virtually 100 % of MM patients [1], but only 20–30 % of patients have demonstrable abnormal karyotypes. This observation is related to the low proliferative index observed in most of the patients. Thus, abnormal cytogenetics can be considered a surrogate of high proliferative index and patients presenting with demonstrable karyotypic abnormalities will have a shorter OS. Due to the limitation of conventional cytogenetics, the International Myeloma Working Group (IMWG) recommendation is for baseline testing with either cytoplasmic immunoglobulin-enhanced fluorescent in situ hybridization (FISH) or FISH carried out on the nuclei from purified plasma cells. The probes used should be able to detect t(4;14)(p16;q32), t(14;16)(q32;q23), t(11; 14)(q13;q32), 17p13 deletions, chromosome 13 deletion, ploidy category, and chromosome 1 abnormalities.
In general, MM can be broadly categorised into two genetic categories: hyperdiploid MM and non-hyperdiploid MM.

Hyperdiploid MM

Hyperdiploid MM is diagnosed by FISH in about 50 % of the cases [24] and is generally associated with a better outcome. These patients’ clonal cells harbour multiple trisomies of the odd-numbered chromosomes, with the exception of chromosome 13. Hyperdiploidy has been observed in MGUS [5, 6] and is, therefore, considered to be an initiating pathogenetic event.
Rarely, hyperdiploidy is associated with high-risk cytogenetics such as IgH translocations, del(17p) and +1q abnormality. It is still unclear if this favourable prognostic impact is lost in such cases as recent studies give us conflicting results, with two studies demonstrating a negative impact on survival [7, 8], whereas Kumar et al. [9] found that the presence of even one chromosomal trisomy was able to ameliorate the adverse effect of the t(4;14), t(14;16), and t(14;20).

Non-hyperdiploid MM

Non-hyperdiploid MM is characterised by a relative high prevalence of IgH chromosome translocations (>85 %) [3, 4]. Most of these patients, with the exception of those with a t(11;14)(q13;q32) and t(6;14) translocation, have an aggressive disease course, characterised by rapid relapse and shorter survival [1018].

IgH Translocations

Patients with translocations t(4;14)(p16;q32) and t(14;16)(q32;q23),seen in 15 and 5 % of cases, respectively, are associated with high-risk MM [12, 14]. The t(4;14)(p16;q32) is associated with FGFR3 and MMSET dysregulation [19], while t(14;16)(q32;q23) is associated with increased c-maf expression [20]. These patients also have a higher frequency of additional genetic abnormalities that are associated with poor prognosis such as chromosome 13 and 14 deletion, chromosome 17 abnormalities, chromosome 1q amplication, and 1p deletion [24]. Another maf gene is b-maf, the expression of which is highly correlated with t(14;20)(q32;q12), a translocation occurring in about 1–2 % of all MM cases and strongly associated with poor prognosis [2123].

17p Deletion [del(17p)]

These patients present with more aggressive disease, extramedullary disease (EMD), and central nervous system involvement [14, 24]. During the disease course, any primary genetic subtype can acquire del(17p), thereby conferring a worse prognosis.

Other Cytogenetic Abnormalities

Initially, chromosome 13 deletion was thought to negatively affect both event-free(EFS) and overall survival (OS) but more recent studies have suggested that its prognostic value is actually related to other genetic changes, especially t(4;14) and del(17p) [10]. Hypodiploidy detected by karyotype has also been identified as an adverse prognostic factor [2529].

Gene Expression Profiling

Two large studies using gene expression profiling have been reported. These two reports describe two different sets of genes that identify poor-risk patient populations. One of the two models—the University of Arkansas for Medical Sciences (UAMS) 70-gene model—has identified genes that are involved in cell cycle regulation, angiogenesis, cell adhesion, cell migration, and proliferation. About 30 % of the informative genes are mapped to chromosome 1 [17]. In the other model—the Intergroupe Francophone du Myélome (IFM) 15-gene model—high-risk patients were enriched in genes controlling proliferation, whereas low-risk patients were enriched in hyperdiploid karyotypes [30].
Multivariate analysis comparing the GEP70-gene model with the 15-gene model revealed that the GEP70-gene model was significant in all datasets tested, but the 15-gene model was significant in bortezomib trials only [31]. It is interesting to note that the two models do not have a single common gene, reflecting either differences in techniques, the treatment used to define the patient population and/or redundancy in the genes and pathways that control growth, proliferation, and survival.

Other Significant Molecular Abnormalities

Amplification of 1q21 increases the risk of MM progression and incidence of the amplification is higher in relapsed than in newly diagnosed MM [32, 33]. Gains of 1q are associated with upregulation of genes involved in intracellular protein transport, including COPA and ARF1 and downregulation of three genes involved in protein translation (RPLP2, RPL21, and FAU) [34]. Cases with 1q gains also showed significantly modulated expression of genes involved in ER stress-induced responses. The unfolded protein response promotes cell survival by reducing the accumulation of aberrantly folded proteins through translation arrest, production of chaperone proteins, and increased degradation [35].
In the MRC Myeloma IX trial, DNA methylation profiling was carried out. A few locally regulated genes were identified—CD38, GPX3, NCAM1/CD56, PDK4, RASD1, RBP1, SPARC,and TGFBI [36]. Hypermethylation of these genes was associated with significantly shorter OS, independent of age, International Staging System score (ISS), and adverse cytogenetics [36]. Although not commonly seen in MM, BRAF V600E is associated with significantly shorter OS in patients and increased aggressiveness [3739].

Tumour Burden

Indicators of tumour burden such as LDH, plasma cell proliferative rate, extramedullary disease, and disease aggressiveness play an important role in MM prognosis. Assessment of disease burden includes imaging modalities and laboratory measurements.

Disease Presentation

Initial presentation of the disease also has prognostic significance. A high stage at diagnosis, plasma cell leukaemia and extramedullary disease (EMD) affect response to therapy and survival.
Extramedullary Disease
Extramedullary disease portends poor prognosis. For the patients enrolled into the Total Therapy protocols, a 5-year OS was 31 % for those with EMD documented by baseline PET/CT, compared with 59 % for those who did not have EMD [40]. Several cytogenetic abnormalities, especially those associated with proliferation and aggressive disease, such as t(14;16), t(14;20), and del(17p), has been associated with EMD [41, 42]. High-risk molecular subtypes such as BRAF V600E is up to four times more frequent in EMD compared to marrow disease [37]. As plasmacytomas are a manifestation of EMD, patients presenting with this condition generally have a shorter OS [4346].
Plasma Cell Leukaemia
Plasma cell leukaemia (PCL) represents a unique subset of patients with MM and can be associated with del(13q) and del(17p) [47] and MYC translocation [42]. These patients tend to do poorly compared to other MM patients as demonstrated in a large series, where the median OS was only 16 months despite treatment with novel agents [48].

Imaging

The type of imaging modality is more important than ever with the updated IMWG criteria for diagnosis of MM. The new criteria consider any bone lesions seen on skeletal survey (SKS), CT, and PET/CT significant enough to make a diagnosis of MM [49••].
Whole Body MRI (WBMRI)
In several studies including one of the largest using MRI, additional skeletal abnormalities were detected by MRI compared to SKS [50, 51]. This is because MRI has high sensitivity for the early detection of marrow infiltration by myeloma cells. IMWG regards more than one focal lesion of a diameter >5 mm as diagnostic of symptomatic MM [52].
To be of prognostic significance, seven or more lesions are needed before MRI can be an independent prognostic factor of survival [51]. The authors reported that the resolution of focal lesions correlated with superior OS [51]. The number of focal lesions on WBMRI significantly correlates with and predicts OS in patients in the context of ASCT [53]. WBMRI has been shown to predict risk of relapse after ASCT if residual disease is detected [53, 54].
Studies have shown that patients with a diffuse pattern of infiltration had features of more advanced disease and poorer OS [5557]. It is thought that the diffuse pattern was associated with the extent of angiogenesis and increased microvessel density in the BM. Moulopoulos et al. demonstrated that there was an improvement of survival in patients who had a diffuse pattern of bone marrow involvement on MRI if they were treated with novel agents [58]. In contrast to these findings, a study from the University of Arkansas involving MM patients treated with a tandem ASCT-based protocol found that heterogeneity of diffuse bone marrow signal did not emerge as an independent adverse feature for OS on multivariate analysis.
PET/CT
PET/CT can detect MM osteolytic lesion with a sensitivity of 80–90 % and specificity 80–100 %. PET/CT can detect more lesions than MRI, with the exception of spinal disease. FDG PET/CT can predict response to ASCT, in that a negative FDG PET/CT after autologous SCT can be considered predictive for non-relapse and, conversely, a short time-to-relapse can be anticipated in patients in whom significantly increased metabolic activity is still persistent after therapy [5961]. The presence of more than three FDG-avid focal lesions is an independent parameter associated with decreased OS and EFS [62]. Many have reported that the presence of extramedullary lesions and the increased value of SUVmax predicted poor prognosis with shorter OS [40, 61, 63].

Laboratory Measurement

Lactate Dehydrogenase
Serum lactate dehydrogenase (LDH) was not included in the ISS staging system because it was not found to be an independent prognostic factor on multivariate analysis [64]. This finding was challenged by other studies which have found that LDH does provide a convenient and dependable prognostic indicator in MM, and a high LDH is associated with worse outcomes [65, 66].
Measurement of Plasma Cell Proliferation
Plasma cell labelling index (PCLI) has a prognostic value, but this technique is expensive and can only be performed in a few laboratories. Cell morphology has also been considered a prognostic variable in MM but the drawback of this method is the considerable interobserver variability. A test looking at the fractal dimension on chromatin of May-Grunwald–Giemsa-stained myeloma cells has shown prognostic significance, where a higher fractal dimension is associated with poorer survival [67]. This test would need to be validated but may be a simple, inexpensive alternative.
Free Light Chain
Measurement of serum free light chain (sFLC) has long been established as part of the diagnostic workup of plasma cell dyscrasias. The quantity of the free light chain measured and the kappa to lambda ratio (rFLC) has prognostic significance from MGUS to active MM. One of the new diagnostic criteria for MM is a serum involved/uninvolved free light chain ratio of 100 or greater, provided the absolute level of the involved free light chain is at least 100 mg/L [49••]. Solitary plasmacytoma also demonstrated that an abnormal rFLC and a persistent serum M protein level of ≥0.5 g/dL was an additional risk factor for progression to MM. Using these two variables, a risk-stratification model was constructed with low-risk, intermediate-risk and high-risk groups having a 5-year progression rate of 13, 26 and 62 %, respectively (P < 0.001) [68, 69]. Even in active MM, several studies have shown that baseline sFLC is prognostic for survival in patients with newly diagnosed active myeloma [6873]. Van Rhee et al. have also demonstrated that among 301 patients enrolled to receive Total Therapy III, those with the highest levels of FLC—greater than 750 mg/l, had the poorest outcomes [71].
There is a concern that the widely used ISS staging system may not prognosticate accurately for ethnic Chinese patients, as there is less accuracy for determination of disease stages in Chinese patients with MM especially between stages II and III [74]. A recent study confirms that both sFLC and rFLC has prognostic significance in MM for Chinese patients as seen in European patients [68, 69, 7173, 75]. The authors of the study suggested combining rFLC to the current ISS, using rFLC (<0.04 or >25), serum β2M (>3.5 mg/L) and serum albumin (<35 g/L) as three adverse risk factors. Patients in the low risk group (no factor, the low-intermediate risk group (one factor), the high-intermediate risk group (two factors) and the high-risk group (three factors) had a median OS of NR, NR, 24 months and 13 months,respectively (p < 0.05) [75].
Heavy Light Chain Specific Immunoglobulin Ratios
Traditional screening tests for monoclonal gammopathies are serum protein electrophoresis (SPE) with scanning densitometry and/or immunofixation electrophoresis (IFE) together with serum free light chain immunoassays. While SPE is a simple, low cost test, it is not very sensitive and quantification of proteins at low concentrations (1–3 g/L) is inaccurate. Improved sensitivity is achieved with IFE, but IFE is a non-quantitative assay. A new method was recently developed and validated for the separate quantification of the kappa- and lambda-bounded amounts of circulating IgG and IgA [7678]. This was achieved by developing antibodies targeting unique junctional epitopes between the heavy and light chains of each Ig molecule. Studies have shown that involved Ig (iHLC) is related to parameters of disease activity, severity, and tumour burden, including anaemia, sFLCR, and extensive bone marrow infiltration [7780]. Koulieris et al. also concluded that the ratio of involved/uninvolved Ig (HLCR) correlated with time to treatment and, most importantly, with outcome; thus, patients with ‘high’ HLCR had a significantly shorter OS [79]. The impact of HLCR on PFS and OS has since been confirmed by others [77, 78, 81, 82]. Ludwig et al. concluded that patients who are deemed having achieved CR by conventional testing (SPEP, NEPH and IFE) may still have residual disease by HLC testing. Vice versa, conversion of a normal to an abnormal HLCR indicated evolving relapse, weeks or months before relapse became evident by conventional methods [78].

Host Factors

Prognosis in MM also depends on host factors (age, performance status, frailty, comorbidities and response to therapy) which will be discussed below.

Age

Increasing age has been purported to be a poor prognostic factor in MM. In a recent study of 682 newly diagnosed patients, 23 % were 80 years or older [83]. These patients had worse performance status, advanced ISS, and a median OS of 22 months, with 14 % dying within 2 months of initial therapy [83]. In a meta-analysis of 1435 patients 65 years or older enrolled in four European phase III trials that included thalidomide and/or bortezomib, the risk for death was increased in patients 75 years or older (HR, 1.44; 95 % CI, 1.2–1.72; P < 0.001) in multivariate analysis [84]. The inferior outcome seen may be that older patients have more comorbidities, their MM is more aggressive biologically, or they tolerate therapies more poorly. The observation by San Miguel et al. that the initial performance status had a highly significant relationship with survival can be generalised to reflect age as it tends to be associated with performance status [85]. However, even in the ‘younger’ myeloma patients (60–65 years of age), there is evidence that age affects the outcome with more ISS II and III in the ‘older’ group (60–65 years) vs. the ‘younger’ group (<60 years) [86]. There was no difference in high-risk cytogenetics between the two groups but the ‘older’ group had a higher beta 2 microglobulin (β2m) which in turn, accounted for higher ISS [86]. It is currently unclear why the β2m levels are increased with age.

Frailty Score

In cancer patients, frailty and Comprehensive Geriatric Assessment (CGA) are being incorporated to guide treatment decisions [87] Frailty is a state of cumulative decline in many physiological systems, resulting in a diminished resistance to stressors, such as cancer and its treatment [8890]. It must not be confused with age and performance status. Frailty profiles were associated with a shorter OS, regardless of staging and treatment administered. The higher risk of death for disease progression was related with lower dose intensity as a consequence of higher rate of drug discontinuation and/or dose reduction and a higher cumulative incidence of nonhematologic toxicities and drug discontinuation [84].
By combining the frailty score with the established ISS, the 3-year OS rate was 55 % in the frail-ISS 3 group, and 94 % in the fit-ISS 1 group. The combination of these two independent parameters significantly improved the prognostic value of each and is an important strategy for predicting outcome. Although not evidence based, treatment can be tailored to frailty scores, with fit patients receiving intensive therapies including stem cell transplant, intermediate-fitness patients treated with doublet treatments or less intense triplets and frail patients receiving gentler, reduced-dose therapy or even a palliative/supportive treatment.

Response to Treatment

There are data to show that depth of response is associated with better outcome. Most studies showed that achieving good response is an important surrogate for improved OS [9193]. In fact, in the Spanish [92] and Italian [94] trials, the achievement of complete remission (CR) has a significant impact on survival when compared with very good partial response only. The Arkansas group and our own data have demonstrated that it is not only the achievement of CR but the ability to sustain the CR that is important for overall survival [9597]. Loss of CR within a short time of achieving it (early relapse) is associated with a worse outcome compared to those who never achieved a CR. IMWG and EBMT harmonised their response criteria and a new category of stringent CR (sCR) was established as a way to measure the quality of response, although there was prognostic significance. sCR is defined as having normal rFLC, absence of clonal cells in the bone marrow, negative IFE on the serum and urine, disappearance of any soft tissue plasmacytomas, and <5 % plasma cells in the bone marrow [98]. Despite its wide use as a clinical endpoint, only one study [99] has reported a benefit of sCR over CR,whereas other studies suggested that the rFLC do not provide additional prognostication [100102]. The issue is further complicated by oligoclonal bands that develop post ASCT that causes a false positive sFLC result [103].
A more robust measure of quality remission is the achievement of immunophenotypic remission [104, 105] and molecular CR [106, 107] with multiple studies demonstrating association with longer survival. The prognostic implications of achieving molecular CR have been discussed elsewhere in this journal. A detection of disease by MRI [53] or PET/CT [61] after treatment is associated with a high risk of relapse. This suggests that response to treatment as well as depth of response is an important prognostic factor. The best treatment to bring a patient into deep remission as well as the best measure of response is still unclear. It is conceivable that a combination of measures would be needed to define the deepest response.

Current Prognostication Systems

There are many known prognostic factors in MM and several different risk stratification systems have been described in MM. Most of these risk stratification systems utilised prognostic factors that were described before the advent of novel agents and it is questionable if they still hold true as novel agents are able to overcome the high-risk features defined by one or more of these prognostic factors. They also do not help in making therapeutic decisions.

Durie-Salmon Staging System and International Staging System

The Durie-Salmon (DS) staging system was developed 30 years ago to provide a practical way to measure MM tumour burden [108]. It measures calcium, renal function, haemoglobin, and the number of lytic lesions. This system is predictive of clinical outcome after standard-dose chemotherapy, but one of the limitations of the DS staging system is the interobserver variability in the number of lytic lesions seen on a skeletal survey. Moreover, the DS staging system predates magnetic resonance imaging (MRI) and positron emission tomography-computed tomography (PET/CT), which are much more sensitive in detecting osseous disease. In 2003, the IMWG introduced the newer Durie-Salmon PLUS (DS+) staging system, which takes into account MRI and PET/CT findings [109, 110]. Comparing the two systems, the DS staging system is based on radiographic findings and clinical parameters, whereas the DS+ staging system relies mainly on modern imaging findings and does not include haemoglobin and immunoglobin levels and may be at a disadvantage in terms of clinical correlation.
In an effort to develop a more clinically relevant, objective and feasible staging system for MM, the International Staging System (ISS) was proposed. A number of statistically significant parameters were analysed and serum β2M and serum albumin were the most consistent, broadly applicable prognostic factors correlated with survival duration. Other parameters (low platelets, high serum creatinine, age, low haemoglobin, calcium, lactate dehydrogenase, and bone marrow plasma cell percentage) had a significant hazard ratio but they identified small patient subsets compared to β2M and serum albumin [64]. Cytogenetic abnormalities that can be identified by FISH have prognostic implications but practical application of this technique has been hampered by lack of standardisation, costs, and restricted availability (at the time of publication of the ISS staging system). Furthermore, to enable the staging system to be widely adopted, the consensus was to use parameters that are routinely used and easily available.
The ISS segregates patients with MM into stages I, II, and III which have an associated median OS of 62, 44 and 29 months, respectively [64]. The ISS is based on the measurement of serum albumin and β2M levels. However, cutoff levels remained a matter of controversy because renal failure could elevate β2M levels even in patients with low tumour burden. Therefore, this system cannot consistently provide a good estimate of tumour burden.

Development of a Better Stratification System

Combining iFISH with other prognostic factors has been examined by several groups. Previously published results by the Intergroupe Francophone du Myélome [10, 11], one of the German groups, [111] and Mayo Clinic mSMART model [112] showed that using iFISH produced a good separation of the patients, based on outcome. Avet-Loiseau et al. combined ISS with cytogenetic abnormalities to improve stratification of MM patients [11]. There was a good differentiation of patients with the 4-year PFS 39, 20, and 11 % and OS 71, 45, and 33 % for ISS-iFISH groups I, II and III, respectively. Although iFISH would be offered widely, cost may be a limiting factor for widespread use.

Revised-ISS

As mentioned previously, LDH is a relevant biomarker in MM. Recent data from three separate myeloma trials has shown that patients with t(4;14) and/or del(17p) in addition to ISS stage III and/or high levels of LDH are at high risk of progression-related death despite modern treatment strategies [113••]. These findings demonstrate that the prognostic accuracy of the ISS can be enhanced through the incorporation of both biochemical and genetic parameters. The new stratification system proposed the revised international staging system (R-ISS): R-ISS I, (ISS stage I, no high-risk CA [del(17p) and/or t(4;14) and/or t(14;16)], and normal LDH); R-ISS III, (ISS stage III and high-risk CA or high LDH level; and R-ISS II, including all the other possible combinations (Table 1). At a median follow-up of 46 months, the 5-year OS rate was 82 % in the R-ISS I, 62 % in the R-ISS II, and 40 % in the R-ISS III groups; the 5-year PFS rates were 55, 36, and 24 %, respectively. The limitations of this new stratification system are the exclusion of chromosome 1 abnormalities as a prognostic parameter, no interlaboratory standardisation of FISH analysis, and heterogeneous cutoff levels for LDH and cost.
Table 1
Revised International Staging System (R-ISS) [99]
Stage
ISS
High-risk CA [del(17p) and/or t(4;14) and/or t(14;16)]
LDH level
R-ISS I
I
None
Normal
R-ISS II
I
Presenta
Higha
 
II
None
Normal
 
II
Presenta
Higha
 
III
None
Normal
R-ISS III
III
Presenta
Higha
aif either positive

Conclusion

There are many variables that affect risk stratification. Table 2 summarises the main factors.
Table 2
Summary of main prognostic variables
Tumour factors
Cytogenetic abnormalities
•Stratifies patients into high risk or standard risk
Tumour burden
•Extramedullary disease and plasma cell leukaemia presentation confers worse prognosis
MRI
•>1 lesion of a diameter >5 mm is diagnostic of MM
•Diffuse pattern of infiltration had features of more advanced disease and poorer OS
PET/CT
•Good for detecting EMD
•EMD and increased SUVmax associated with poorer OS
LDH
•Newly incorporated into the revised ISS as a independent prognostic factor
Host factors
Frailty score
•Frailty profiles are associated with poorer OS
•Can help tailor treatment
Response to treatment
•Depth of response is associated with better outcome
•Immunophenotypic and molecular CR are more robust in predicting OS
It is important to predict outcome accurately by using simple and easily applicable tools. This has implications in patient counselling and risk-adapted treatment. Furthermore, development of a uniform risk stratification system allows better comparison of patient groups across different trials. The current ISS staging system is outdated as the Revised-ISS can prognosticate MM patients more accurately.
The challenge would be to incorporate other important prognostic factors into the current risk stratification systems. It is clearly demonstrated that GEP is important in risk stratification of MM but the lack of universal availability impedes its inclusion in the current staging system. Treatment response and further modification of the stratification systems based on response attained will also be important to evaluate. Eventually, establishing a staging system that is able to prognosticate accurately and enable a risk-adapted treatment strategy that is modifiable based on response achieved would be an important goal to advance our current management approaches in MM.

Compliance with Ethical Standards

Conflict of Interest

The authors declare that they have no competing interests.

Human and Animal Rights and Informed Consent

This article does not contain any studies with human or animal subjects performed by any of the authors.
Literature
1.
Avet-Loiseau H, Li C, Magrangeas F, Gouraud W, Charbonnel C, Harousseau J-L, et al. Prognostic significance of copy-number alterations in multiple myeloma. J Clin Oncol Off J Am Soc Clin Oncol. 2009;27(27):4585–90. doi:10.​1200/​jco.​2008.​20.​6136.CrossRef
2.
Debes-Marun CS, Dewald GW, Bryant S, Picken E, Santana-Dávila R, González-Paz N, et al. Chromosome abnormalities clustering and its implications for pathogenesis and prognosis in myeloma. Leukemia. 2003;17(2):427–36. doi:10.​1038/​sj.​leu.​2402797.PubMedCrossRef
3.
Fonseca R, Debes-Marun CS, Picken EB, Dewald GW, Bryant SC, Winkler JM, et al. The recurrent IgH translocations are highly associated with nonhyperdiploid variant multiple myeloma. Blood. 2003;102(7):2562–7. doi:10.​1182/​blood-2003-02-0493.PubMedCrossRef
4.
Smadja NV, Leroux D, Soulier J, Dumont S, Arnould C, Taviaux S, et al. Further cytogenetic characterization of multiple myeloma confirms that 14q32 translocations are a very rare event in hyperdiploid cases. Genes Chromosomes Cancer. 2003;38(3):234–9. doi:10.​1002/​gcc.​10275.PubMedCrossRef
5.
Brousseau M, Leleu X, Gerard J, Gastinne T, Godon A, Genevieve F, et al. Hyperdiploidy is a common finding in monoclonal gammopathy of undetermined significance and monosomy 13 is restricted to these hyperdiploid patients. Clin Cancer Res. 2007;13(20):6026–31. doi:10.​1158/​1078-0432.​ccr-07-0031.PubMedCrossRef
6.
Chng WJ, Van Wier SA, Ahmann GJ, Winkler JM, Jalal SM, Bergsagel PL, et al. A validated FISH trisomy index demonstrates the hyperdiploid and nonhyperdiploid dichotomy in MGUS. Blood. 2005;106(6):2156–61. doi:10.​1182/​blood-2005-02-0761.PubMedPubMedCentralCrossRef
7.
Chng WJ, Santana-Davila R, Van Wier SA, Ahmann GJ, Jalal SM, Bergsagel PL, et al. Prognostic factors for hyperdiploid-myeloma: effects of chromosome 13 deletions and IgH translocations. Leukemia. 2006;20(5):807–13. doi:10.​1038/​sj.​leu.​2404172.PubMedCrossRef
8.
Pawlyn C, Melchor L, Murison A, Wardell CP, Brioli A, Boyle EM, et al. Coexistent hyperdiploidy does not abrogate poor prognosis in myeloma with adverse cytogenetics and may precede IGH translocations. Blood. 2015;125(5):831–40. doi:10.​1182/​blood-2014-07-584268.PubMedPubMedCentralCrossRef
9.
Kumar S, Fonseca R, Ketterling RP, Dispenzieri A, Lacy MQ, Gertz MA, et al. Trisomies in multiple myeloma: impact on survival in patients with high-risk cytogenetics. Blood. 2012;119(9):2100–5. doi:10.​1182/​blood-2011-11-390658.PubMedPubMedCentralCrossRef
10.
Avet-Loiseau H, Attal M, Moreau P, Charbonnel C, Garban F, Hulin C, et al. Genetic abnormalities and survival in multiple myeloma: the experience of the Intergroupe Francophone du Myélome. Blood. 2007;109(8):3489–95. doi:10.​1182/​blood-2006-08-040410.PubMedCrossRef
11.
Avet-Loiseau H, Durie BGM, Cavo M, Attal M, Gutierrez N, Haessler J, et al. Combining fluorescent in situ hybridization data with ISS staging improves risk assessment in myeloma: an International Myeloma Working Group collaborative project. Leukemia. 2013;27(3):711–7. doi:10.​1038/​leu.​2012.​282.PubMedPubMedCentralCrossRef
12.
Avet-Loiseau H, Facon T, Grosbois B, Magrangeas F, Rapp M-J, Harousseau J-L, et al. Oncogenesis of multiple myeloma: 14q32 and 13q chromosomal abnormalities are not randomly distributed, but correlate with natural history, immunological features, and clinical presentation. Blood. 2002;99(6):2185–91.PubMedCrossRef
13.
Chang H, Sloan S, Li D, Keith Stewart A. Multiple myeloma involving central nervous system: high frequency of chromosome 17p13.1 (p53) deletions. Br J Haematol. 2004;127(3):280–4. doi:10.​1111/​j.​1365-2141.​2004.​05199.​x.PubMedCrossRef
14.
Fonseca R, Blood E, Rue M, Harrington D, Oken MM, Kyle RA, et al. Clinical and biologic implications of recurrent genomic aberrations in myeloma. Blood. 2003;101(11):4569–75. doi:10.​1182/​blood-2002-10-3017.PubMedCrossRef
15.
Keats JJ, Reiman T, Maxwell CA, Taylor BJ, Larratt LM, Mant MJ, et al. In multiple myeloma, t(4;14)(p16;q32) is an adverse prognostic factor irrespective of FGFR3 expression. Blood. 2003;101(4):1520–9. doi:10.​1182/​blood-2002-06-1675.PubMedCrossRef
16.
Pineda-Roman M, Zangari M, van Rhee F, Anaissie E, Szymonifka J, Hoering A, et al. VTD combination therapy with bortezomib-thalidomide-dexamethasone is highly effective in advanced and refractory multiple myeloma. Leukemia. 2008;22(7):1419–27. doi:10.​1038/​leu.​2008.​99.PubMedPubMedCentralCrossRef
17.
Shaughnessy Jr JD, Zhan F, Burington BE, Huang Y, Colla S, Hanamura I, et al. A validated gene expression model of high-risk multiple myeloma is defined by deregulated expression of genes mapping to chromosome 1. Blood. 2007;109(6):2276–84. doi:10.​1182/​blood-2006-07-038430.PubMedCrossRef
18.
Nemec P, Zemanova Z, Kuglik P, Michalova K, Tajtlova J, Kaisarova P, et al. Complex karyotype and translocation t(4;14) define patients with high-risk newly diagnosed multiple myeloma: results of CMG2002 trial. Leuk Lymphoma. 2012;53(5):920–7. doi:10.​3109/​10428194.​2011.​634042.PubMedCrossRef
19.
Chesi M, Nardini E, Brents LA, Schröck E, Ried T, Kuehl WM, 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(3):260–4. doi:10.​1038/​ng0797-260.PubMedPubMedCentralCrossRef
20.
Chesi M, Bergsagel PL, Shonukan OO, Martelli ML, Brents LA, Chen T, et al. Frequent dysregulation of the c-maf proto-oncogene at 16q23 by translocation to an Ig locus in multiple myeloma. Blood. 1998;91(12):4457–63.PubMed
21.
Boersma-Vreugdenhil GR, Kuipers J, Van Stralen E, Peeters T, Michaux L, Hagemeijer A, et al. The recurrent translocation t(14;20)(q32;q12) in multiple myeloma results in aberrant expression of MAFB: a molecular and genetic analysis of the chromosomal breakpoint. Br J Haematol. 2004;126(3):355–63. doi:10.​1111/​j.​1365-2141.​2004.​05050.​x.PubMedCrossRef
22.
Ross FM, Chiecchio L, Dagrada G, Protheroe RK, Stockley DM, Harrison CJ, et al. The t(14;20) is a poor prognostic factor in myeloma but is associated with long-term stable disease in monoclonal gammopathies of undetermined significance. Haematologica. 2010;95(7):1221–5. doi:10.​3324/​haematol.​2009.​016329.PubMedPubMedCentralCrossRef
23.
Vekemans MC, Lemmens H, Delforge M, Doyen C, Pierre P, Demuynck H, et al. The t(14;20)(q32;q12): a rare cytogenetic change in multiple myeloma associated with poor outcome. Br J Haematol. 2010;149(6):901–4. doi:10.​1111/​j.​1365-2141.​2010.​08113.​x.PubMedCrossRef
24.
Chang H, Qi C, Yi Q-L, Reece D, Stewart AK. p53 gene deletion detected by fluorescence in situ hybridization is an adverse prognostic factor for patients with multiple myeloma following autologous stem cell transplantation. Blood. 2005;105(1):358–60. doi:10.​1182/​blood-2004-04-1363.PubMedCrossRef
25.
Calasanz MJ, Cigudosa JC, Odero MD, García-Foncillas J, Marín J, Ardanaz MT, et al. Hypodiploidy and 22q11 rearrangements at diagnosis are associated with poor prognosis in patients with multiple myeloma. Br J Haematol. 1997;98(2):418–25.PubMedCrossRef
26.
Fassas AB-T, Spencer T, Sawyer J, Zangari M, Lee C-K, Anaissie E, et al. Both hypodiploidy and deletion of chromosome 13 independently confer poor prognosis in multiple myeloma. Br J Haematol. 2002;118(4):1041–7.PubMedCrossRef
27.
Smadja NV, Bastard C, Brigaudeau C, Leroux D, Fruchart C, Hématologique GFdC. Hypodiploidy is a major prognostic factor in multiple myeloma. Blood. 2001;98(7):2229–38.PubMedCrossRef
28.
Van Wier S, Braggio E, Baker A, Ahmann G, Levy J, Carpten JD, et al. Hypodiploid multiple myeloma is characterized by more aggressive molecular markers than non-hyperdiploid multiple myeloma. Haematologica. 2013;98(10):1586–92. doi:10.​3324/​haematol.​2012.​081083.PubMedPubMedCentralCrossRef
29.
Oh S, Koo DH, Kwon MJ, Kim K, Suh C, Min CK, et al. Chromosome 13 deletion and hypodiploidy on conventional cytogenetics are robust prognostic factors in Korean multiple myeloma patients: web-based multicenter registry study. Ann Hematol. 2014;93(8):1353–61. doi:10.​1007/​s00277-014-2057-5.PubMedCrossRef
30.
Decaux O, Lode L, Magrangeas F, Charbonnel C, Gouraud W, Jezequel P, et al. Prediction of survival in multiple myeloma based on gene expression profiles reveals cell cycle and chromosomal instability signatures in high-risk patients and hyperdiploid signatures in low-risk patients: a study of the Intergroupe Francophone du Myelome. J Clin Oncol Off J Am Soc Clin Oncol. 2008;26(29):4798–805. doi:10.​1200/​jco.​2007.​13.​8545.CrossRef
31.
Johnson SK, Heuck CJ, Albino AP, Qu P, Zhang Q, Barlogie B, et al. The use of molecular-based risk stratification and pharmacogenomics for outcome prediction and personalized therapeutic management of multiple myeloma. Int J Hematol. 2011;94(4):321–33. doi:10.​1007/​s12185-011-0948-y.PubMedPubMedCentralCrossRef
32.
Hanamura I, Stewart JP, Huang Y, Zhan F, Santra M, Sawyer JR, et al. Frequent gain of chromosome band 1q21 in plasma-cell dyscrasias detected by fluorescence in situ hybridization: incidence increases from MGUS to relapsed myeloma and is related to prognosis and disease progression following tandem stem-cell transplantation. Blood. 2006;108(5):1724–32. doi:10.​1182/​blood-2006-03-009910.PubMedPubMedCentralCrossRef
33.
Rosiñol L, Carrió A, Bladé J, Queralt R, Aymerich M, Cibeira MT, et al. Comparative genomic hybridisation identifies two variants of smoldering multiple myeloma. Br J Haematol. 2005;130(5):729–32. doi:10.​1111/​j.​1365-2141.​2005.​05673.​x.PubMedCrossRef
34.
Carew JS, Nawrocki ST, Krupnik YV, Dunner K, McConkey DJ, Keating MJ, et al. Targeting endoplasmic reticulum protein transport: a novel strategy to kill malignant B cells and overcome fludarabine resistance in CLL. Blood. 2006;107(1):222–31. doi:10.​1182/​blood-2005-05-1923.PubMedPubMedCentralCrossRef
35.
Fabris S, Ronchetti D, Agnelli L, Baldini L, Morabito F, Bicciato S, et al. Transcriptional features of multiple myeloma patients with chromosome 1q gain. Leukemia. 2007;21(5):1113–6. doi:10.​1038/​sj.​leu.​2404616.PubMed
36.
Kaiser MF, Johnson DC, Wu P, Walker BA, Brioli A, Mirabella F, et al. Global methylation analysis identifies prognostically important epigenetically inactivated tumor suppressor genes in multiple myeloma. Blood. 2013;122(2):219–26. doi:10.​1182/​blood-2013-03-487884.PubMedPubMedCentralCrossRef
37.
Andrulis M, Lehners N, Capper D, Penzel R, Heining C, Huellein J, et al. Targeting the BRAF V600E mutation in multiple myeloma. Canc Discov. 2013;3(8):862–9. doi:10.​1158/​2159-8290.​CD-13-0014.CrossRef
38.
Bohn OL, Hsu K, Hyman DM, Pignataro DS, Giralt S, Teruya-Feldstein J. BRAF V600E mutation and clonal evolution in a patient with relapsed refractory myeloma with plasmablastic differentiation. Clin Lymphoma Myeloma Leuk. 2014;14(2):e65–8. doi:10.​1016/​j.​clml.​2013.​12.​003.PubMedCrossRef
39.
Sharman JP, Chmielecki J, Morosini D, Palmer GA, Ross JS, Stephens PJ, et al. Vemurafenib response in 2 patients with posttransplant refractory BRAF V600E-mutated multiple myeloma. Clin Lymphoma Myeloma Leuk. 2014;14(5):e161–3. doi:10.​1016/​j.​clml.​2014.​06.​004.PubMedCrossRef
40.
Usmani SZ, Mitchell A, Waheed S, Crowley J, Hoering A, Petty N, et al. Prognostic implications of serial 18-fluoro-deoxyglucose emission tomography in multiple myeloma treated with total therapy 3. Blood. 2013;121(10):1819–23. doi:10.​1182/​blood-2012-08-451690.PubMedPubMedCentralCrossRef
41.
Lopez-Anglada L, Gutierrez NC, Garcia JL, Mateos MV, Flores T, San Miguel JF. P53 deletion may drive the clinical evolution and treatment response in multiple myeloma. Eur J Haematol. 2010;84(4):359–61. doi:10.​1111/​j.​1600-0609.​2009.​01399.​x.PubMedCrossRef
42.
Tiedemann RE, Gonzalez-Paz N, Kyle RA, Santana-Davila R, Price-Troska T, Van Wier SA, et al. Genetic aberrations and survival in plasma cell leukemia. Leukemia. 2008;22(5):1044–52. doi:10.​1038/​leu.​2008.​4.PubMedPubMedCentralCrossRef
43.
Fassas AB, Ward S, Muwalla F, Van Hemert R, Schluterman K, Harik S, et al. Myeloma of the central nervous system: strong association with unfavorable chromosomal abnormalities and other high-risk disease features. Leuk Lymphoma. 2004;45(2):291–300.PubMedCrossRef
44.
Rasche L, Bernard C, Topp MS, Kapp M, Duell J, Wesemeier C, et al. Features of extramedullary myeloma relapse: high proliferation, minimal marrow involvement, adverse cytogenetics: a retrospective single-center study of 24 cases. Ann Hematol. 2012;91(7):1031–7. doi:10.​1007/​s00277-012-1414-5.PubMedCrossRef
45.
Terpos E, Rezvani K, Basu S, Milne AE, Rose PE, Scott GL, et al. Plasmacytoma relapses in the absence of systemic progression post-high-dose therapy for multiple myeloma. Eur J Haematol. 2005;75(5):376–83. doi:10.​1111/​j.​1600-0609.​2005.​00531.​x.PubMedCrossRef
46.
Varettoni M, Corso A, Pica G, Mangiacavalli S, Pascutto C, Lazzarino M. Incidence, presenting features and outcome of extramedullary disease in multiple myeloma: a longitudinal study on 1003 consecutive patients. Ann Oncol. 2010;21(2):325–30. doi:10.​1093/​annonc/​mdp329.PubMedCrossRef
47.
Chang H, Sloan S, Li D, Patterson B. Genomic aberrations in plasma cell leukemia shown by interphase fluorescence in situ hybridization. Cancer Genet Cytogenet. 2005;156(2):150–3. doi:10.​1016/​j.​cancergencyto.​2004.​05.​004.PubMedCrossRef
48.
Avet-Loiseau H, Roussel M, Campion L, Leleu X, Marit G, Jardel H, et al. Cytogenetic and therapeutic characterization of primary plasma cell leukemia: the IFM experience. Leukemia. 2012;26(1):158–9. doi:10.​1038/​leu.​2011.​176.PubMedCrossRef
49.••
Rajkumar SV, Dimopoulos MA, Palumbo A, Blade J, Merlini G, Mateos MV, et al. International Myeloma Working Group updated criteria for the diagnosis of multiple myeloma. Lancet Oncol. 2014;15(12):e538–48. doi:10.​1016/​s1470-2045(14)70442-5. This article is an update of the criteria for the diagnosis of MM. The revised criteria makes the diagnosis of MM earlier. PubMedCrossRef
50.
Song IC, Kim JN, Choi YS, Ryu H, Lee MW, Lee HJ, et al. Diagnostic and prognostic implications of spine magnetic resonance imaging at diagnosis in patients with multiple myeloma. Cancer Res Treat. 2015;47(3):465–72. doi:10.​4143/​crt.​2014.​010.PubMedPubMedCentralCrossRef
51.
Walker R, Barlogie B, Haessler J, Tricot G, Anaissie E, Shaughnessy JD, et al. Magnetic resonance imaging in multiple myeloma: diagnostic and clinical implications. J Clin Oncol Off J Am Soc Clin Oncol. 2007;25(9):1121–8. doi:10.​1200/​jco.​2006.​08.​5803.CrossRef
52.
Dimopoulos MA, Hillengass J, Usmani S, Zamagni E, Lentzsch S, Davies FE, et al. Role of magnetic resonance imaging in the management of patients with multiple myeloma: a consensus statement. J Clin Oncol Off J Am Soc Clin Oncol. 2015;33(6):657–64. doi:10.​1200/​JCO.​2014.​57.​9961.CrossRef
53.
Hillengass J, Ayyaz S, Kilk K, Weber MA, Hielscher T, Shah R, et al. Changes in magnetic resonance imaging before and after autologous stem cell transplantation correlate with response and survival in multiple myeloma. Haematologica. 2012;97(11):1757–60. doi:10.​3324/​haematol.​2012.​065359.PubMedPubMedCentralCrossRef
54.
Bannas P, Hentschel HB, Bley TA, Treszl A, Eulenburg C, Derlin T, et al. Diagnostic performance of whole-body MRI for the detection of persistent or relapsing disease in multiple myeloma after stem cell transplantation. Eur Radiol. 2012;22(9):2007–12. doi:10.​1007/​s00330-012-2445-y.PubMedCrossRef
55.
Baur A, Stäbler A, Nagel D, Lamerz R, Bartl R, Hiller E, et al. Magnetic resonance imaging as a supplement for the clinical staging system of Durie and Salmon? Cancer. 2002;95(6):1334–45. doi:10.​1002/​cncr.​10818.PubMedCrossRef
56.
Moulopoulos LA, Gika D, Anagnostopoulos A, Delasalle K, Weber D, Alexanian R, et al. Prognostic significance of magnetic resonance imaging of bone marrow in previously untreated patients with multiple myeloma. Ann Oncol. 2005;16(11):1824–8. doi:10.​1093/​annonc/​mdi362.PubMedCrossRef
57.
Stäbler A, Baur A, Bartl R, Munker R, Lamerz R, Reiser MF. Contrast enhancement and quantitative signal analysis in MR imaging of multiple myeloma: assessment of focal and diffuse growth patterns in marrow correlated with biopsies and survival rates. AJR Am J Roentgenol. 1996;167(4):1029–36. doi:10.​2214/​ajr.​167.​4.​8819407.PubMedCrossRef
58.
Moulopoulos LA, Dimopoulos MA, Kastritis E, Christoulas D, Gkotzamanidou M, Roussou M, et al. Diffuse pattern of bone marrow involvement on magnetic resonance imaging is associated with high risk cytogenetics and poor outcome in newly diagnosed, symptomatic patients with multiple myeloma: a single center experience on 228 patients. Am J Hematol. 2012;87(9):861–4. doi:10.​1002/​ajh.​23258.PubMedCrossRef
59.
Derlin T, Peldschus K, Munster S, Bannas P, Herrmann J, Stubig T, et al. Comparative diagnostic performance of (1)(8)F-FDG PET/CT versus whole-body MRI for determination of remission status in multiple myeloma after stem cell transplantation. Eur Radiol. 2013;23(2):570–8. doi:10.​1007/​s00330-012-2600-5.PubMedCrossRef
60.
Nanni C, Zamagni E, Celli M, Caroli P, Ambrosini V, Tacchetti P, et al. The value of 18F-FDG PET/CT after autologous stem cell transplantation (ASCT) in patients affected by multiple myeloma (MM): experience with 77 patients. Clin Nucl Med. 2013;38(2):e74–9. doi:10.​1097/​RLU.​0b013e318266cee2​.PubMedCrossRef
61.
Zamagni E, Patriarca F, Nanni C, Zannetti B, Englaro E, Pezzi A, et al. Prognostic relevance of 18-F FDG PET/CT in newly diagnosed multiple myeloma patients treated with up-front autologous transplantation. Blood. 2011;118(23):5989–95. doi:10.​1182/​blood-2011-06-361386.PubMedCrossRef
62.
Bartel TB, Haessler J, Brown TLY, Shaughnessy JD, van Rhee F, Anaissie E, et al. F18-fluorodeoxyglucose positron emission tomography in the context of other imaging techniques and prognostic factors in multiple myeloma. Blood. 2009;114(10):2068–76. doi:10.​1182/​blood-2009-03-213280.PubMedPubMedCentralCrossRef
63.
Haznedar R, Aki SZ, Akdemir OU, Ozkurt ZN, Ceneli O, Yagci M, et al. Value of 18F-fluorodeoxyglucose uptake in positron emission tomography/computed tomography in predicting survival in multiple myeloma. Eur J Nucl Med Mol Imaging. 2011;38(6):1046–53. doi:10.​1007/​s00259-011-1738-8.PubMedCrossRef
64.
Greipp PR, San Miguel J, Durie BG, Crowley JJ, Barlogie B, Blade J, et al. International staging system for multiple myeloma. J Clin Oncol. 2005;23(15):3412–20. doi:10.​1200/​jco.​2005.​04.​242.PubMedCrossRef
65.
Gkotzamanidou M, Kastritis E, Gavriatopoulou MR, Nikitas N, Gika D, Mparmparousi D, et al. Increased serum lactate dehydrongenase should be included among the variables that define very-high-risk multiple myeloma. Clin Lymphoma Myeloma Leuk. 2011;11(5):409–13. doi:10.​1016/​j.​clml.​2011.​07.​001.PubMedCrossRef
66.
Terpos E, Katodritou E, Roussou M, Pouli A, Michalis E, Delimpasi S, et al. High serum lactate dehydrogenase adds prognostic value to the international myeloma staging system even in the era of novel agents. Eur J Haematol. 2010;85(2):114–9. doi:10.​1111/​j.​1600-0609.​2010.​01466.​x.PubMed
67.
Ferro DP, Falconi MA, Adam RL, Ortega MM, Lima CP, de Souza CA, et al. Fractal characteristics of May-Grunwald-Giemsa stained chromatin are independent prognostic factors for survival in multiple myeloma. PLoS One. 2011;6(6), e20706. doi:10.​1371/​journal.​pone.​0020706.PubMedPubMedCentralCrossRef
68.
Dispenzieri A, Zhang L, Katzmann JA, Snyder M, Blood E, Degoey R, et al. Appraisal of immunoglobulin free light chain as a marker of response. Blood. 2008;111(10):4908–15. doi:10.​1182/​blood-2008-02-138602.PubMedPubMedCentralCrossRef
69.
Snozek CL, Katzmann JA, Kyle RA, Dispenzieri A, Larson DR, Therneau TM, et al. Prognostic value of the serum free light chain ratio in newly diagnosed myeloma: proposed incorporation into the international staging system. Leukemia. 2008;22(10):1933–7. doi:10.​1038/​leu.​2008.​171.PubMedPubMedCentralCrossRef
70.
Kyrtsonis MC, Vassilakopoulos TP, Kafasi N, Sachanas S, Tzenou T, Papadogiannis A, et al. Prognostic value of serum free light chain ratio at diagnosis in multiple myeloma. Br J Haematol. 2007;137(3):240–3. doi:10.​1111/​j.​1365-2141.​2007.​06561.​x.PubMedCrossRef
71.
van Rhee F, Bolejack V, Hollmig K, Pineda-Roman M, Anaissie E, Epstein J, et al. High serum-free light chain levels and their rapid reduction in response to therapy define an aggressive multiple myeloma subtype with poor prognosis. Blood. 2007;110(3):827–32. doi:10.​1182/​blood-2007-01-067728.PubMedPubMedCentralCrossRef
72.
Khoriaty R, Hussein MA, Faiman B, Kelly M, Kalaycio M, Baz R. Prediction of response and progression in multiple myeloma with serum free light chains assay: corroboration of the serum free light chain response definitions. Clin Lymphoma Myeloma Leuk. 2010;10(1):E10–3. doi:10.​3816/​CLML.​2010.​n.​010.PubMedCrossRef
73.
Sthaneshwar P, Nadarajan V, Maniam JA, Nordin N, Gin Gin G. Serum free light chains: diagnostic and prognostic value in multiple myeloma. Clin Chem Lab Med. 2009;47(9):1101–7. doi:10.​1515/​cclm.​2009.​260.PubMedCrossRef
74.
Deng SH, Xu Y, Mai YJ, Wang YF, Zhao YZ, Zou DH, et al. Analysis of the international staging system of multiple myeloma and its comparison with the DS and IFM staging system in 122 Chinese patients. Zhonghua xue ye xue za zhi = Zhonghua xueyexue zazhi. 2008;29(4):217–21.PubMed
75.
Xu Y, Sui W, Deng S, An G, Wang Y, Xie Z, et al. Further stratification of patients with multiple myeloma by International Staging System in combination with ratio of serum free kappa to lambda light chains. Leuk Lymphoma. 2013;54(1):123–32. doi:10.​3109/​10428194.​2012.​704033.PubMedCrossRef
76.
Bradwell AR, Harding SJ, Fourrier NJ, Wallis GLF, Drayson MT, Carr-Smith HD, et al. Assessment of monoclonal gammopathies by nephelometric measurement of individual immunoglobulin kappa/lambda ratios. Clin Chem. 2009;55(9):1646–55. doi:10.​1373/​clinchem.​2009.​123828.PubMedCrossRef
77.
Bradwell A, Harding S, Fourrier N, Mathiot C, Attal M, Moreau P, et al. Prognostic utility of intact immunoglobulin Ig‘kappa/Ig’lambda ratios in multiple myeloma patients. Leukemia. 2013;27(1):202–7. doi:10.​1038/​leu.​2012.​159.PubMedPubMedCentralCrossRef
78.
Ludwig H, Milosavljevic D, Zojer N, Faint JM, Bradwell AR, Hübl W, et al. Immunoglobulin heavy/light chain ratios improve paraprotein detection and monitoring, identify residual disease and correlate with survival in multiple myeloma patients. Leukemia. 2013;27(1):213–9. doi:10.​1038/​leu.​2012.​197.PubMedPubMedCentralCrossRef
79.
Koulieris E, Panayiotidis P, Harding SJ, Kafasi N, Maltezas D, Bartzis V, et al. Ratio of involved/uninvolved immunoglobulin quantification by Hevylite™ assay: clinical and prognostic impact in multiple myeloma. Exp Hematol Oncol. 2012;1(1):9. doi:10.​1186/​2162-3619-1-9.PubMedPubMedCentralCrossRef
80.
Batinić J, Perić Z, Šegulja D, Last J, Prijić S, Dubravčić K, et al. Immunoglobulin heavy/light chain analysis enhances the detection of residual disease and monitoring of multiple myeloma patients. Croat Med J. 2015;56(3):263–71.PubMedPubMedCentralCrossRef
81.
Beaumont-Epinette MP, Moreau C, Besnard S, Latute F, Collet N, Sebillot M, et al. Heavy/light chain specific immunoglobulin ratios provides no additional information than serum proteins electrophoresis and immunofixation for the diagnosis and the follow-up of intact immunoglobulin multiple myeloma patients. Pathol Biol. 2015;63(4-5):215–21. doi:10.​1016/​j.​patbio.​2015.​06.​001.PubMedCrossRef
82.
Tovar N, Fernández de Larrea C, Elena M, Cibeira MT, Aróstegui JI, Rosiñol L, et al. Prognostic impact of serum immunoglobulin heavy/light chain ratio in patients with multiple myeloma in complete remission after autologous stem cell transplantation. Biol Blood Marrow Transplant. 2012;18(7):1076–9. doi:10.​1016/​j.​bbmt.​2012.​03.​004.PubMedCrossRef
83.
Dimopoulos MA, Kastritis E, Delimpasi S, Katodritou E, Hatzimichael E, Kyrtsonis M-C, et al. Multiple myeloma in octogenarians: clinical features and outcome in the novel agent era. Eur J Haematol. 2012;89(1):10–5. doi:10.​1111/​j.​1600-0609.​2012.​01784.​x.PubMedCrossRef
84.
Bringhen S, Mateos MV, Zweegman S, Larocca A, Falcone AP, Oriol A, et al. Age and organ damage correlate with poor survival in myeloma patients: meta-analysis of 1435 individual patient data from 4 randomized trials. Haematologica. 2013;98(6):980–7. doi:10.​3324/​haematol.​2012.​075051.PubMedPubMedCentralCrossRef
85.
San Miguel JF, Sànchez J, Gonzalez M. Prognostic factors and classification in multiple myeloma. Br J Cancer. 1989;59(1):113–8.PubMedPubMedCentralCrossRef
86.
Chretien ML, Hebraud B, Cances-Lauwers V, Hulin C, Marit G, Leleu X, et al. Age is a prognostic factor even among patients with multiple myeloma younger than 66 years treated with high-dose melphalan: the IFM experience on 2316 patients. Haematologica. 2014;99(7):1236–8. doi:10.​3324/​haematol.​2013.​098608.PubMedPubMedCentralCrossRef
87.
Hamaker ME, Jonker JM, de Rooij SE, Vos AG, Smorenburg CH, van Munster BC. Frailty screening methods for predicting outcome of a comprehensive geriatric assessment in elderly patients with cancer: a systematic review. Lancet Oncol. 2012;13(10):e437–44. doi:10.​1016/​s1470-2045(12)70259-0.PubMedCrossRef
88.
Clegg A, Young J, Iliffe S, Rikkert MO, Rockwood K. Frailty in elderly people. Lancet. 2013;381(9868):752–62. doi:10.​1016/​s0140-6736(12)62167-9.PubMedCrossRef
89.
Fried LP, Tangen CM, Walston J, Newman AB, Hirsch C, Gottdiener J, et al. Frailty in older adults: evidence for a phenotype. J Gerontol A Biol Sci Med Sci. 2001;56(3):M146–56.PubMedCrossRef
90.
Slaets JPJ. Vulnerability in the elderly: frailty. Med Clin North Am. 2006;90(4):593–601. doi:10.​1016/​j.​mcna.​2006.​05.​008.PubMedCrossRef
91.
Harousseau J-L, Avet-Loiseau H, Attal M, Charbonnel C, Garban F, Hulin C, et al. Achievement of at least very good partial response is a simple and robust prognostic factor in patients with multiple myeloma treated with high-dose therapy: long-term analysis of the IFM 99-02 and 99-04 Trials. J Clin Oncol Off J Am Soc Clin Oncol. 2009;27(34):5720–6. doi:10.​1200/​jco.​2008.​21.​1060.CrossRef
92.
Lahuerta JJ, Mateos MV, Martínez-López J, Rosiñol L, Sureda A, de la Rubia J, et al. Influence of pre- and post-transplantation responses on outcome of patients with multiple myeloma: sequential improvement of response and achievement of complete response are associated with longer survival. J Clin Oncol Off J Am Soc Clin Oncol. 2008;26(35):5775–82. doi:10.​1200/​jco.​2008.​17.​9721.CrossRef
93.
Moreau P, Attal M, Pégourié B, Planche L, Hulin C, Facon T, et al. Achievement of VGPR to induction therapy is an important prognostic factor for longer PFS in the IFM 2005-01 trial. Blood. 2011;117(11):3041–4. doi:10.​1182/​blood-2010-08-300863.PubMedCrossRef
94.
Gay F, Larocca A, Wijermans P, Cavallo F, Rossi D, Schaafsma R, et al. Complete response correlates with long-term progression-free and overall survival in elderly myeloma treated with novel agents: analysis of 1175 patients. Blood. 2011;117(11):3025–31. doi:10.​1182/​blood-2010-09-307645.PubMedCrossRef
95.
Barlogie B, Anaissie E, Haessler J, van Rhee F, Pineda-Roman M, Hollmig K, et al. Complete remission sustained 3 years from treatment initiation is a powerful surrogate for extended survival in multiple myeloma. Cancer. 2008;113(2):355–9. doi:10.​1002/​cncr.​23546.PubMedCrossRef
96.
Hoering A, Crowley J, Shaughnessy JD, Hollmig K, Alsayed Y, Szymonifka J, et al. Complete remission in multiple myeloma examined as time-dependent variable in terms of both onset and duration in Total Therapy protocols. Blood. 2009;114(7):1299–305. doi:10.​1182/​blood-2009-03-211953.PubMedPubMedCentralCrossRef
97.
Ong S, Widanalage S, Chen YX, Ooi MG, Surendran S, Koh LP, et al. Early relapse post autologous transplant is a stronger predictor of survival compared to pretreatment patient factors in the novel agent era - Analysis of Singapore Multiple Myeloma Working Group. 2015.
98.
Durie B, Harousseau J, Miguel J, Blade J, Barlogie B, Anderson K, et al. International uniform response criteria for multiple myeloma. Leukemia. 2006;20(9):1467–73.PubMedCrossRef
99.
Kapoor P, Kumar SK, Dispenzieri A, Lacy MQ, Buadi F, Dingli D, et al. Importance of achieving stringent complete response after autologous stem-cell transplantation in multiple myeloma. J Clin Oncol Off J Am Soc Clin Oncol. 2013;31(36):4529–35. doi:10.​1200/​jco.​2013.​49.​0086.CrossRef
100.
Kroger N, Asenova S, Gerritzen A, Bacher U, Zander A. Questionable role of free light chain assay ratio to determine stringent complete remission in multiple myeloma patients. Blood. 2010;115(16):3413–4. doi:10.​1182/​blood-2010-01-261677. author reply 3414-3415.PubMedCrossRef
101.
Paiva B, Martinez-Lopez J, Vidriales MB, Mateos MV, Montalban MA, Fernandez-Redondo E, et al. Comparison of immunofixation, serum free light chain, and immunophenotyping for response evaluation and prognostication in multiple myeloma. J Clin Oncol Off J Am Soc Clin Oncol. 2011;29(12):1627–33. doi:10.​1200/​jco.​2010.​33.​1967.CrossRef
102.
Martinez-Lopez J, Paiva B, Lopez-Anglada L, Mateos MV, Cedena T, Vidriales MB, et al. Critical analysis of the stringent complete response in multiple myeloma: contribution of sFLC and bone marrow clonality. Blood. 2015;126(7):858–62. doi:10.​1182/​blood-2015-04-638742.PubMedPubMedCentralCrossRef
103.
de Larrea CF, Cibeira MT, Elena M, Arostegui JI, Rosinol L, Rovira M, et al. Abnormal serum free light chain ratio in patients with multiple myeloma in complete remission has strong association with the presence of oligoclonal bands: implications for stringent complete remission definition. Blood. 2009;114(24):4954–6. doi:10.​1182/​blood-2009-06-224832.PubMedCrossRef
104.
Paiva B, Gutiérrez NC, Rosiñol L, Vídriales M-B, Montalbán M-Á, Martínez-López J, et al. High-risk cytogenetics and persistent minimal residual disease by multiparameter flow cytometry predict unsustained complete response after autologous stem cell transplantation in multiple myeloma. Blood. 2012;119(3):687–91. doi:10.​1182/​blood-2011-07-370460.PubMedCrossRef
105.
Paiva B, Vidriales M-B, Cerveró J, Mateo G, Pérez JJ, Montalbán MA, et al. Multiparameter flow cytometric remission is the most relevant prognostic factor for multiple myeloma patients who undergo autologous stem cell transplantation. Blood. 2008;112(10):4017–23. doi:10.​1182/​blood-2008-05-159624.PubMedPubMedCentralCrossRef
106.
Kröger N, Badbaran A, Zabelina T, Ayuk F, Wolschke C, Alchalby H, et al. Impact of high-risk cytogenetics and achievement of molecular remission on long-term freedom from disease after autologous-allogeneic tandem transplantation in patients with multiple myeloma. Biol Blood Marrow Transplant. 2013;19(3):398–404. doi:10.​1016/​j.​bbmt.​2012.​10.​008.PubMedCrossRef
107.
Ladetto M, Pagliano G, Ferrero S, Cavallo F, Drandi D, Santo L, et al. Major tumor shrinking and persistent molecular remissions after consolidation with bortezomib, thalidomide, and dexamethasone in patients with autografted myeloma. J Clin Oncol. 2010;28(12):2077–84. doi:10.​1200/​jco.​2009.​23.​7172.PubMedCrossRef
108.
Durie BG, Salmon SE. A clinical staging system for multiple myeloma. Correlation of measured myeloma cell mass with presenting clinical features, response to treatment, and survival. Cancer. 1975;36(3):842–54.PubMedCrossRef
109.
Kimura N, Tokunaga C, Dalal S, Richardson C, Yoshino K, Hara K, et al. A possible linkage between AMP-activated protein kinase (AMPK) and mammalian target of rapamycin (mTOR) signalling pathway. Genes Cells. 2003;8(1):65–79.PubMedCrossRef
110.
Durie BG, Kyle RA, Belch A, Bensinger W, Blade J, Boccadoro M, et al. Myeloma management guidelines: a consensus report from the Scientific Advisors of the International Myeloma Foundation. Hematol J. 2003;4(6):379–98. doi:10.​1038/​sj.​thj.​6200312.PubMedCrossRef
111.
Neben K, Lokhorst HM, Jauch A, Bertsch U, Hielscher T, van der Holt B, et al. Administration of bortezomib before and after autologous stem cell transplantation improves outcome in multiple myeloma patients with deletion 17p. Blood. 2012;119(4):940–8. doi:10.​1182/​blood-2011-09-379164.PubMedCrossRef
112.
Kapoor P, Fonseca R, Rajkumar SV, Sinha S, Gertz MA, Stewart AK, et al. Evidence for cytogenetic and fluorescence in situ hybridization risk stratification of newly diagnosed multiple myeloma in the era of novel therapie. Mayo Clin Proc. 2010;85(6):532–7. doi:10.​4065/​mcp.​2009.​0677.PubMedPubMedCentralCrossRef
113.••
Moreau P, Cavo M, Sonneveld P, Rosinol L, Attal M, Pezzi A, et al. Combination of international scoring system 3, high lactate dehydrogenase, and t(4;14) and/or del(17p) identifies patients with multiple myeloma (MM) treated with front-line autologous stem-cell transplantation at high risk of early MM progression-related death. J Clin Oncol. 2014;32(20):2173–80. doi:10.​1200/​jco.​2013.​53.​0329. This article explains the new revised International Staging System.PubMedCrossRef