Genetic Factors: Hereditary Cancer Predisposition Syndromes


Summary of Key Points

  • The discovery of inherited mutations of genes associated with increased risk for cancer provides important clinical opportunities for early detection and prevention of common and rare forms of human malignancies.

  • Syndromes of cancer predisposition often involve multiple organ systems, affect paired organs with bilateral or multifocal tumors, and have onset at an earlier age compared with nonfamilial tumors. The diagnosis of particular cancer predisposition syndromes can usually be confirmed with molecular genetic testing of patients who have hereditary malignancies. Genetic testing can then be extended to relatives as a predictive test to guide their preventive management.

  • Medical, surgical, and radiation oncologists, genetic counselors, and allied professionals are playing a leading role in the integration of genetic testing into the practice of preventive oncology. Recently, genomic analysis has been applied to tumors to discover targets for therapy. Because tumor genomic analysis will also include a comparison with the germline, the need for genetic counseling for both cancer and noncancer disease risks will be increased. This chapter reviews both common and more recently described familial cancer syndromes, with an emphasis on the clinical application of cancer genetic and genomic analysis in the management of patients who have or are at risk for cancer.

During the past decade, the availability of clinical testing for inherited mutations of cancer predisposition genes has had a major impact on the practice of clinical oncology. As these genes were identified and characterized, guidelines for the responsible clinical translation of this information were developed by medical and surgical subspecialty societies, such as the American Society of Clinical Oncology (updated in 2016), the American Association for Cancer Research Pediatric Cancer Working Group, and other organizations. These guidelines emphasized that in the process of offering a predictive genetic test to a patient or family that is affected by cancer, the provider and the individual who is being tested must be prepared to deal with all the medical, psychological, and social consequences of a positive, negative, or ambiguous result. These guidelines define the form and content of genetic counseling as a component of cancer risk assessment and management.

A selected set of syndromes of cancer predisposition is reviewed in this chapter ( Tables 13.1 and 13.2 ). Detailed discussions of breast, colon, gastric, gynecologic, pancreatic, prostate, and renal cancer susceptibility are found in the chapters that discuss these tumors. Whether offered by a physician, genetic counselor, or other health care professional, genetic testing for inherited cancer risk requires careful informed consent. The elements of informed consent for genetic testing are summarized in Box 13.1 . With the advent of genomic screening of tumors, gene panel testing, and genomic screening in the germline, these elements of informed consent must also include the eventuality of “incidental” findings associated with risk of cancer, noncancer diseases, and nonpaternity.

Table 13.1
Pathologic Hallmarks of Selected Cancer Predisposition Syndromes
Neoplastic or Preneoplastic Pathologic Conditions Associated Genetic Syndromes (Gene)
Atypical rhabdoid tumor, schwannomatosis Rhabdoid tumor predisposition syndrome (SMARCB1)
Wilms tumor Li-Fraumeni syndrome (TP53), (WT1),(BWS), Simpson-Golabi-Behmel syndrome, Beckwith-Wiedemann syndrome, Fanconi anemia compound heterozygotes (see Fanconi anemia)
Ductal breast cancer with estrogen receptor, progesterone receptor, and HER2-neu negativity (“triple negative”) Hereditary breast and ovarian cancer syndrome (BRCA1)
Lobular breast cancer Hereditary diffuse gastric cancer syndrome (CDH1)
High-grade serous (papillary serous) ovarian cancer Hereditary breast and ovarian cancer syndrome (see Table 13.2 )
Hepatoblastoma Familial adenomatous polyposis (APC), Beckwith-Wiedemann syndrome, Simpson-Golabi-Behmel syndrome, Sotos syndrome
Hereditary paraganglioma-pheochromocytoma Hereditary paraganglioma-pheochromocytoma syndromes (SDHB, SDHC, SDHD, SDHAF2)
Hypodiploid leukemia Li-Fraumeni syndrome (TP53)
Clear cell or endometrioid ovarian cancers Lynch syndrome (MLH1, MSH2, MSH6, PMS2, EPCAM)
Ovarian sex cord tumors with annular tubules (SCTATs) Peutz-Jeghers syndrome (STK11)
Clear cell or endometrioid endometrial cancers Lynch syndrome (MLH1, MSH2, MSH6, PMS2, EPCAM)
Uterine leiomyomas Hereditary leiomyoma renal cell carcinoma syndrome (FH)
Adenoma malignum of the cervix Peutz-Jeghers syndrome (STK11)
Colorectal cancer with tumor infiltrating lymphocytes, Crohnlike lymphocytic reaction, mucinous or signet ring cells, medullary growth pattern Lynch syndrome (MLH1, MSH2, MSH6, PMS2, EPCAM)
Lauren diffuse type, signet ring cell, or linitus plastica gastric cancer Hereditary diffuse gastric cancer syndrome (CDH1)
Leiomyosarcoma Li-Fraumeni syndrome ( TP53 ), retinoblastoma (RB1)
Hamartomatous gastrointestinal polyps Peutz-Jeghers syndrome (STK11) ; juvenile polyposis syndrome (BMPR1A, SMAD4) ; Cowden syndrome (PTEN)
Hypermutable tumors Constitutional mismatch repair deficiency, Lynch syndrome (MLH1 MSH2 MSH6 PMS2 EPCAM)
Paget breast tumor Li-Fraumeni syndrome (TP53)
Pancreatic neuroendocrine tumors Multiple endocrine neoplasia type 1 (MEN1) ; von Hippel-Lindau disease (VHL)
Renal papillary carcinoma type I Hereditary papillary renal cancer syndrome (MET)
Renal papillary carcinoma type II, collecting duct, tubulopapillary Hereditary leiomyoma renal cell carcinoma syndrome (FH)
Renal clear cell carcinoma von Hippel-Lindau disease (VHL)
Renal oncocytoma Birt-Hogg-Dube syndrome (FLCN ); tuberous sclerosis complex (TSC1)
Renal angiomyolipoma Tuberous sclerosis complex (TSC1)
Renal chromophobe Birt-Hogg-Dube syndrome (FLCN)
Renal medullary cancer Sickle cell trait
Sertoli cell tumors of the testes Peutz-Jeghers syndrome (STK11) ;
Small cell carcinoma of the ovary–hypercalcemic type (SCCOHT) Rhabdoid tumor predisposition syndrome (SMARCA4)
Medullary thyroid cancer Multiple endocrine neoplasia type 2 (RET)
Meningioma Neurofibromatosis type 2 (NF2), Gorlin syndrome (PTCH1)
Cribriform-morula variant of papillary thyroid carcinoma Familial adenomatous polyposis (APC)
Follicular thyroid cancer Cowden syndrome (PTEN)
Medulloblastoma Turcot syndrome variant of familial adenomatous polyposis (APC) ; Gorlin syndrome (PTCH1), Fanconi anemia ( FANC complex genes),
Glioblastoma Turcot syndrome variant of Lynch syndrome (MLH1, MSH2, MSH6, nPMS2)
Basal cell carcinoma Xeroderma pigmentosum ( XPA,B,C,D; DDB2, ERCC4, ERCC5, POLH) ; Gorlin syndrome (PTCH1)
Schwannomatosis Neurofibromatosis type 2; rhabdoid predisposition syndrome (SMARCB1, LZTR1)
Sebaceous adenoma, sebaceous carcinoma, keratoacanthoma Muir Torre syndrome variant of Lynch syndrome (MLH1, MSH2)
Fibrofolliculomas Birt-Hogg-Dube syndrome (FLCN)
Facial tricholemmomas Cowden syndrome (PTEN)

Table 13.2
Cancer Susceptibility Genes, Syndromes, and Management Considerations
The majority of childhood onset conditions are not listed. Please refer to Lindor NM, McMaster ML, Lindor CJ. Concise handbook of familial cancer susceptibility syndromes, 2nd ed. J Natl Cancer Inst Monogr. 2008;(38):1-93. PMID:18559331
Genes and Details Syndrome Penetrance Associated Neoplasms Management Considerations Other Notes
APC
Tumor suppressor
Dominant inheritance
Familial adenomatous polyposis (FAP)
Attenuated FAP (AFAP)
Variant: Turcot syndrome
High Colorectal lesions (adenomatous polyposis and cancer)
Gastric and duodenal or small bowel lesions (adenomatous polyps and cancer)
Desmoid tumors
Thyroid (typically papillary)
Hepatoblastoma (patients usually ≤5 yr old)
Pancreatic
medulloblastoma (Turcot syndrome)
Colonoscopy and total colectomy based on polyp burden
Upper endoscopy
Consideration of small bowel visualization
Thyroid examination and consideration of ultrasound examination
Consideration of childhood liver palpation, abdominal ultrasound examination, AFP measurement (investigational)
Genotype-phenotype correlations between classic FAP and AFAP: Mutations on the 5′ and 3′ ends of the gene correlate with AFAP
High de novo mutation rate (~20%–25%)
Extracolonic features: Congenital hypertrophy of the retinal pigment epithelium (CHRPE); osteomas, supernumerary teeth; desmoids, epidermoid cysts; gastric fundic gland polyps
Particularly cribriform-morula variant of papillary thyroid carcinoma
Chemoprevention not a replacement for colonoscopy and colectomy
APC*I1307K (c.3920T>A) Moderate Colorectal Colonoscopy based on family history or similar to first-degree relative risk Ashkenazi Jewish founder mutation
Not associated with polyposis
ATM Ataxia telangiectasia (AT) Moderate (monoallelic/ heterozygous carriers)
Rare: High (biallelic)
Monoallelic carriers:
Breast, possibly pancreas
Biallelic carriers:
Leukemia, lymphoma
Monoallelic carriers: Breast screening with annual mammogram and breast magnetic resonance imaging (MRI) starting at age 40 yr or per family history a
Biallelic carriers: AT specialist for multidisciplinary management: screening complete blood count (CBC) with differential; consider bone marrow biopsy, AFP measurement
Immunology: Monitoring immunoglobulin levels per immunologist recommendation
Dermatology: annual skin examination
Pulmonary: Baseline and as-needed pulmonary function tests
Gastroenterology and nutrition: Baseline and as-needed swallowing function and nutritional management
Endocrine: Annual diabetes screen
Neurology: Supportive medications
Orthopedics: Annual scoliosis evaluation
Dental: Biannual examination
Monoallelic carriers: Limited evidence for pancreas or prostate cancer
Risk of autosomal recessive AT condition in biallelic offspring of heterozygous carriers
AT: Childhood cerebellar ataxia, telangiectasias of the conjunctivae, immunodeficiency, sensitivity to radiation
BAP1
Possible tumor suppressor
Dominant inheritance
BAP1 tumor predisposition syndrome Undetermined Atypical Spitz tumors
Uveal melanoma
Malignant mesothelioma
Cutaneous melanoma
Clear cell renal cell carcinoma
Basal cell carcinoma
Eye examinations
Dermatologic examinations
Consideration of abdominal ultrasound examination and/or MRI of the kidneys
Limited evidence for possible other associated tumors: non–small cell lung adenocarcinoma, breast cancer, cholangiocarcinoma, meningioma, neuroendocrine carcinoma
BLM Bloom syndrome Low or undetermined
(monoallelic/ heterozygous carriers)
Rare: High (biallelic)
Biallelic: Myelodysplasia
Variety of epithelial carcinomas (gastrointestinal, genitourinary), lymphoid, hematopoietic, sarcomas, CNS, and others
Biallelic carriers: Bloom syndrome specialist for management:
CBC every 3–4 mo, avoidance of radiation, breast MRI or ultrasound starting at 18 yr of age, annual colonoscopy starting at age 15 yr, renal ultrasound examination at diagnosis and every 3 mo through age 8 yr to assess for Wilms tumor, HPV vaccine per AAP guidelines
Dermatology: Annual skin examination, limit sun exposure
Pulmonary: Pulmonary function tests
Gastroenterology and nutrition: Baseline and as-needed swallowing function evaluation and nutritional management
Endocrine: Annual fasting blood sugar and TSH level
Orthopedics: Annual scoliosis evaluation
Dental: Biannual examination
Monoallelic carriers: Limited evidence for risk of colorectal and breast cancer
Bloom syndrome: Chromosome breakage syndrome with growth deficiency, erythematous facial skin lesion, immunodeficiency, frequent infections
Founder mutation in Ashkenazi Jewish population
BMPR1A and
SMAD4
SMAD4: Tumor suppressor
Dominant inheritance
Juvenile polyposis syndrome High Gastrointestinal lesions (juvenile/ hamartomatous polyps and cancer—colorectal, stomach, small bowel) Colonoscopy based on polyp burden
Upper endoscopy
Surgical management based on polyp burden
Hereditary hemorrhagic telangiectasia (HHT) screening
“Juvenile” refers to polyp histology, not age of onset
SMAD4 mutations associated with HHT
SMAD4 mutations associated with massive gastric polyposis
Limited evidence for association with pancreas
BRCA1 and BRCA2
Tumor suppressor
Dominant inheritance
Hereditary breast and ovarian cancer syndrome (HBOC) High Breast (BRCA1: often triple-negative tumor)
Male breast
Ovarian (epithelial; high-grade serous), fallopian tube, primary peritoneal
Prostate (high Gleason score)
Pancreatic (exocrine)
Breast screening with annual mammogram and breast MRI starting at age 25–30 yr
Male self- and clinical-breast examination
Optional risk-reducing mastectomy
Risk-reducing BSO
PSA measurement and digital rectal examination
Consideration of eligibility and pros and cons of chemoprevention with tamoxifen
Consideration of eligibility for treatment with PARP inhibitors
Investigational-based pancreas screening
Founder mutations in certain populations—Ashkenazi Jewish, Icelandic, and others
Autosomal recessive Fanconi anemia with biallelic germline mutations
BRCA1/2 heterozygous mutations detected in childhood through tumor normal sequencing should be reviewed with clinical genetics team and treating oncology team for importance pertaining to care of relatives and potentially molecularly based therapy
BRIP1
Possible tumor suppressor
Moderate (monoallelic/ heterozygous carriers)
Rare: High (biallelic)
Ovarian (heterozygous carriers) Consider risk-reducing BSO at age 50–55 yr or as per family history b Risk of autosomal recessive Fanconi anemia in biallelic offspring of heterozygous carriers
CDH1
Tumor suppressor
Dominant inheritance
Hereditary diffuse gastric cancer syndrome High Diffuse gastric cancer
Lobular breast cancer
Prophylactic total gastrectomy
Breast screening with annual mammogram and breast MRI
Optional risk-reducing mastectomy
Upper endoscopy not proven to be an effective method of screening or detecting diffuse gastric cancer
Insufficient evidence for colorectal cancer
CDKN2A
Tumor suppressor
CDK4 oncogene
Dominant inheritance
Familial atypical multiple mole melanoma (FAMMM) syndrome High Melanoma and dysplastic nevi
Pancreas
Dermatologic examination
Investigational-based pancreatic screening
Risk of melanoma may be independent of genetic test result
CDKN2a (p16) founder mutation in Netherlands
CHEK2
Tumor suppressor
Dominant inheritance
Moderate Breast
Colon
Breast screening with annual mammogram and breast MRI starting at age 40 yr or per family history a
Colonoscopy based on family history or similar to first-degree relative risk
Genotype-phenotype: Different risks for truncating mutations vs missense mutations
Limited evidence for association with prostate cancer
DICER1
Dominant inheritance
High Pleuropulmonary blastoma (PPB)
Ovarian sex cord-stromal tumors (Sertoli-Leydig cell tumor, juvenile granulose cell tumor, gynandroblastoma)
Cystic nephroma
Thyroid gland neoplasia (cancer, multinodular goiter, adenomas)
No guidelines have been established
Consideration of annual physical examination with targeted systems review and possible imaging
Other features: Ciliary body medulloepithelioma; botryoid-type embryonal rhabdomyosarcoma of the cervix or other sites; nasal chondromesenchymal hamartoma; renal sarcoma; pituitary blastoma; pineoblastoma
Early-onset disease—before age 40 yr
Maternal fetal medicine care when lung cysts are identified prenatally
Papillary or follicular thyroid cancer
GREM1 (SCG5-GREM1)
Dominant inheritance
Hereditary mixed polyposis syndrome High Colorectal lesions (polyps of mixed histologies and cancer) Colonoscopy and surgical management based on polyp burden Only a 40-kb duplication upstream of GREM1 (spanning 3′ end of SCG5 and region upstream of GREM1 ) implicated in disease; identified in Ashkenazi Jewish population
ETV6/CEBPA/PAX5/RUNX1
IKAROS/SAMD9/SAMD9L
Dominant Inheritance
Familial leukemia/MDS High ALL, AML, MDS History and physical examination
Medical history: Prior cytopenias, bleeding history, nononcologic manifestations
Family history:
Review and document types of cancer and leukemia, ages at cancer or leukemia onset
Include history of antecedent cytopenias and/or bleeding
Physical examination:
Signs of leukemia, lymphoma
Other syndrome specific findings, including signs of solid tumors
CBC
Manual differential
Reticulocyte count
Blood smear with morphology
Bone marrow evaluation
Aspirate and biopsy
Morphology
Cytogenetics
Patient and family education about signs and symptoms of cancer, including leukemia
HSCT consultation
Consider HLA typing and genetic testing of potential familial donors
Discuss enrollment in registries or other research studies
Other syndromes predisposing to leukemia include LFS (TP53), CMMRD (MLH1, MSH2, MSH6, PMS2), T21, Bloom (BLM), Nijmegen (NBN), ataxia telangiectasia (ATM), NF1, Noonan: PTPN11, CBL (RAS-activating syndromes), Fanconi anemia (FANCA-E, BRCA, RAD51D [autosomal recessive]), telomere syndromes (TERT, TERC, DKC), severe congenital neutropenia (ELANE, HAX1), Diamond-Blackfan syndrome (RPS19, RPLS, RPL11), Shwachman-Diamond syndrome (SBDS), GATA2, monosomy 7
Fanconi anemia (multiple genes—at least 20)
Recessive inheritance (most genes)
Dominant inheritance ( RAD51 gene)
X-linked inheritance ( FANCB gene)
Fanconi anemia High:
Biallelic for the recessive genes and
monoallelic for RAD51, FANCB, BRCA2, PALB2, BRCA1
Moderate:
Monoallelic
for BRIP1, RAD51C
Myelodysplastic syndrome
Acute myeloid leukemia
Squamous cell carcinoma of head and neck, esophagus, vulva
Genitourinary (cervix, Wilms tumor, neuroblastoma)
Other solid tumors: Liver, brain, skin, breast, gastrointestinal
Associated cancers for BRCA2, PALB2, BRIP1, and RAD51C carriers
Fanconi anemia specialist for management
HSCT
Early detection and surgical management of solid tumors: oral/ otolaryngology/ ear-nose-throat examination; gynecologic examination.
HPV prevention
Management for BRCA2, PALB2, BRIP1, and RAD51C carriers (discussed above and below)
Increased chromosome breakage in lymphocytes tested with diepoxybutane and mitomycin C (does not identify carriers)
Physical abnormalities (not in all patients): Short stature, abnormal skin pigmentation (café au lait or hypopigmentation or hyperpigmentation), skeletal malformations of the upper and lower limbs (thumbs, radii, hands, ulnae), microcephaly, developmental delay, and ophthalmic, genitourinary, cardiac, gastrointestinal, CNS anomalies
Progressive bone marrow failure (not in all patients): Pancytopenia, thrombocytopenia, leucopenia
Extreme toxicities from chemotherapy or radiation
Genes and complement groups:

BRCA2 FA-D1
BRIP1 FA-J
PALB2 FA-N
RAD51 FA-R
RAD51C FA-O
BRCA1 FANCES
XRCC2 FANCU

FH
Tumor suppressor
Dominant inheritance
Hereditary leiomyomatosis and renal cell cancer (HLRCC) High Cutaneous leiomyomata
Uterine leiomyomata
Renal tumors (type 2 papillary, tubule-papillary, collecting duct carcinomas; unilateral, solitary lesions)
Dermatologic and gynecologic examination—evaluate for changes suggestive of leiomyosarcoma
Medication or resection of leiomyomata
Imaging for and surgical management of renal tumors consider starting at age 8
Fumarate hydratase enzyme assay might be helpful in some situations
Risk for uterine leiomyosarcoma is unclear
Biallelic FH mutations cause a recessive disorder known as fumarate hydratase deficiency —metabolic disorder, profound developmental delay, seizures, fumaric aciduria
FLCN
Tumor suppressor
Dominant inheritance
Birt-Hogg-Dube (BHD) syndrome High Cutaneous (fibrofolliculomas, trichodiscomas/ angiofibromas, perifollicular fibromas, acrochordons)
Pulmonary cysts
Renal tumors (oncocytoma, chromophobe, oncocytic hybrid tumors)
No consensus at this time
Dermatologic care
Imaging for and surgical management (nephron-sparing) of renal tumors
Fibrofolliculomas are specific to BHD
Lung cysts are multiple and bilateral
Spontaneous pneumothorax
Renal tumors are multifocal, bilateral, slow growing
Other features: Parotid lesions, oral papules, thyroid lesions
Unclear data regarding colon cancer
Possible genotype-phenotype correlations
KIT (C-Kit)
Oncogene
Dominant inheritance
High Gastrointestinal stromal tumors (GISTs) No consensus at this time
Imaging and endoscopy might be considered
Diffuse interstitial cell of Cajal hyperplasia (ICCH)
Skin hyperpigmentation or hypopigmentation
Mast-cell disorders
Related gene: PDGFRA
MEN1
Tumor suppressor
Dominant inheritance
Multiple endocrine neoplasia type 1 (MEN1) High Gastroenteropancreatic (GEP) tract well- differentiated endocrine tumors
Pituitary tumors (prolactinoma)
Parathyroid tumors
Carcinoid tumors
Adrenocortical tumors
Biochemical testing
Imaging
Surgical management
Various medications
Primary hyperparathyroidism, hypercalcemia, oligomenorrhea or amenorrhea, galactorrhea, Zollinger-Ellison syndrome (gastrinoma), insulinoma, glucagonoma, vasoactive intestinal peptide (VIP)–secreting tumor (VIPomas)
Other features: Skin (angiofibromas, collagenomas), lipomas, CNS lesions (meningioma, ependymoma), leiomyomas
MET (c-Met)
Oncogene
Dominant inheritance
Hereditary papillary renal carcinoma (HPRC) High Papillary renal cancer (multifocal, bilateral, type 1 papillary) Imaging
Surgical management
MLH1, MSH2,
MSH6, PMS2,
EPCAM
Tumor suppressor
Dominant inheritance
Lynch syndrome
(formerly: hereditary non-polyposis colorectal cancer [HNPCC])
Variant: Muir-Torre syndrome
Variant: Turcot syndrome
Variant: Constitutional mismatch repair deficiency syndrome [CMMR-D] (biallelic mutations; recessive inheritance)
High Colorectal
Endometrial
Ovarian (epithelial)
Stomach
Small bowel
Upper urinary tract (renal pelvis, ureter)
Pancreas
Hepatobiliary tract
Sebaceous neoplasms (Muir-Torre syndrome)
Glioblastoma (Turcot syndrome)
Colonoscopy every 1–2 yr starting at age 20–25 yr
Consideration of prophylactic colectomy
Hysterectomy and risk-reducing BSO
Consideration of upper endoscopy
Consideration of urinalysis or urine cytology
Consideration of eligibility for treatment with cancer immunotherapy with checkpoint blockade
Consideration of aspirin for chemoprevention
CMMR-D recessive syndrome
MRI of brain every 6 mo starting at diagnosis
Whole-body MRI starting at age 6 once a year
CBC every 6 mo
Abdominal ultrasound examination every 6 mo
Upper GI endoscopy, VCE, ileocolonoscopy—start at 4–6 yr then every 12 mo
Gynecologic examination, transvaginal ultrasound, urine cytology, dipstick starting at age 20 then every 12 mo
Tumor analyses with immunohistochemical (IHC) staining, microsatellite instability (MSI) analysis, somatic tumor genetic analysis, BRAF V600E somatic analysis, MLH1 promoter hypermethylation analysis
Amsterdam Criteria I and II; Revised Bethesda Guidelines
Recent recommendations for universal screening of colorectal and endometrial cancers
Genotype-phenotype correlations emerging but not significant enough to alter management
Limited evidence for moderate risk of breast and prostate cancer
Autosomal recessive inheritance—biallelic (homozygote or compound heterozygote) with childhood onset of severe CMMR-D syndrome—café au lait macules, solid tumors, hematologic cancers
MUTYH
Recessive inheritance
MUTYH-associated polyposis (MAP) High Colorectal lesions (adenomatous polyposis and cancer; serrated and hyperplastic polyps also described)
Stomach
Duodenal lesions (adenomatous polyps, cancer)
Colonoscopy and total colectomy based on polyp burden
Upper endoscopy
Phenotypically similar to AFAP
Two Northern European founder mutations
Controversy regarding risk in heterozygous carriers (see below)
MUTYH
Monoallelic/ heterozygous carrier
Moderate Colorectal Colonoscopy based on family history or similar to risk with first-degree relative Controversy regarding risk
NBN Nijmegen breakage syndrome (NBS) Moderate
(monoallelic/ heterozygous carriers)
Rare: High (biallelic)
Monoallelic carrier:
Breast
Biallelic carrier:
Lymphoma
Monoallelic carriers: Breast screening with annual mammogram and breast MRI starting age at age 40 yr or as per family history a
Biallelic carriers: NBS specialist for multidisciplinary management
Pediatric autosomal recessive syndrome
Hematology-oncology: History and physical examination, annual CBC, metabolic profile and lactate dehydrogenase, avoid excessive radiation, HPV vaccine per AAP guidelines
Dermatology: Annual skin examination
Pulmonary: Baseline pulmonary function tests with follow-up as needed, aggressive treatment of recurrent infections
Gastroenterology and nutrition: Baseline and as-needed swallowing function evaluation and nutritional management
Endocrine: Monitor growth, assess female patients for ovarian failure
Neurology: Developmental assessment and early intervention if needed
Ophthalmology: Annual examination
Orthopedics: Baseline assessment for anomalies and as needed
Dental: Biannual examination
Heterozygous carrier breast cancer risk; limited evidence for prostate cancer
Slavic founder mutation
Risk of autosomal recessive condition in offspring of heterozygous carriers
(NBS)
NBS: Chromosomal breakage, microcephaly, dysmorphic features, immunodeficiency, lymphomas
NF1
Tumor suppressor
Dominant inheritance
Neurofibromatosis type 1 High (moderate risk for breast cancer) Cutaneous neurofibromas
Plexiform neurofibromas
Optic nerve and other CNS gliomas
Malignant peripheral nerve sheath tumors
Breast (moderate risk)
Others: GIST, leukemia
NF1 specialist for management
Physical examination; ophthalmologic examination; imaging; surgical management
Breast screening with annual mammogram and breast MRI as per family history a
Children with NF1 should have ophthalmic assessments every 6 to 12 mo from birth to 8 yr; one baseline assessment of color vision and visual fields should be undertaken when the child is developmentally able
Assess with history and clinical examination annually for typical signs of MPNST: any nondermal neurofibroma with rapid growth, loss of neurologic function, or increasing pain or change in consistency
Assess for risk of JMML in NF1 in children with juvenile xanthogranulomas
Baseline whole-body MRI should be considered between ages 16 and 20 yr to assess internal tumor burden to determine adult follow-up regimen if signs of MPNST are present
MRI with or without FDG-PET is recommended
Café au lait macules, axillary and inguinal freckling, iris Lisch nodules, learning disabilities, scoliosis, tibial dysplasia, vasculopathy
NF2/SMARCB1/LZTR1/SMARCE1
Dominant inheritance
Neurofibromatosis type 2 High Acoustic neuromas
Schwannomatosis
Meningiomas
Leukemia
Atypical rhabdoid tumor
Annual history and physical examination (including audiology with measurement of pure-tone thresholds and Word Recognition Scores)
Annual (consider twice yearly in first year after diagnosis or signs of rapid growth) brain MRI starting at 10 yr of age; screening may begin earlier in patients with high-risk genotypes or symptomatic diagnoses
If baseline imaging shows no characteristic sites of involvement, reduce frequency of screening to every 2 yr
Protocols should include high-resolution (1- to 3-mm section thickness) imaging through the internal auditory meatus, preferably in at least two orthogonal planes
Surveillance spinal MRI is recommended at 24- to 36-month intervals beginning at 10 yr of age
Whole-body MRI may be performed
PALB2
Tumor suppressor
Dominant inheritance
High Breast
Limited evidence for pancreatic cancer
Breast screening with annual mammogram and breast MRI starting at age 30 yr or per family history a
Optional risk-reducing mastectomy based on family history
Investigational-based pancreas screening
Insufficient evidence for ovarian cancer
Risk of autosomal recessive Fanconi anemia in biallelic offspring of heterozygous carriers
POLD1, POLE
Dominant inheritance
Polymerase proofreading-associated polyposis (PPAP) High Colorectal lesions (polyps and cancer) Colonoscopy and surgical management based on polyp burden Disease-causing mutations occur in the exonuclease domain
Limited data regarding extracolonic cancer risk
PRSS1, SPINK1,
CFTR, CTRC
Various modes of inheritance
Hereditary pancreatitis High (PRSS1)
Variable (SPINK1, CFTR, CTRC)
Pancreas Pancreatitis management Pancreatitis—recurrent, acute, chronic throughout lifetime
More severe disease if biallelic mutations are present
Biallelic (homozygote or compound heterozygote) CFTR mutations associated with cystic fibrosis
PTCH, SUFU
Tumor suppressor
Dominant inheritance
Gorlin syndrome/ nevoid basal cell carcinoma syndrome High Jaw (odontogenic) keratocysts
Basal cell carcinomas
Medulloblastoma (PNET—desmoplastic)
Referral to a specialist for management
Physical examination
Sun exposure avoidance and at least annual dermatology examination
Surgical management
Other features: Congenital skeletal anomalies, cleft lip and palate, cerebral and falx calcifications, macrocephaly with frontal bossing, polydactyly, intellectual disability, lymphomesenteric or pleural cysts, palmar and plantar pits, cardiac fibromas, ovarian fibromas, ocular abnormalities
Related gene: SUFU
PTEN
Tumor suppressor
Dominant inheritance
Cowden syndrome/PTEN
hamartoma tumor syndrome
High Breast
Thyroid (most often follicular)
Endometrial
Colorectal lesions (hamartomas, ganglioneuromas, cancer)
Genitourinary tumors (renal cell carcinoma)
Breast screening with annual mammogram and breast MRI
Optional risk-reducing mastectomy
Thyroid examination and ultrasound
Endometrial screening with random biopsy and ultrasound
Optional hysterectomy
Colonoscopy
Renal imaging
Major and minor criteria (tumors, physical and dermatologic or mucocutaneous findings, developmental disability)
Macrocephaly
Bannayan-Riley-Ruvalcaba syndrome
PTEN-related Proteus syndrome
Autism spectrum disorder
Lhermitte-Duclos disease
RAD51C and RAD51D
Tumor suppressor
Moderate
Rare: High (biallelic RAD51C )
Ovarian Consider risk-reducing BSO at age 50–55 yr or per family history b Risk of autosomal recessive Fanconi anemia in biallelic RAD51C offspring of heterozygous carriers
RB1
Tumor suppressor
Dominant inheritance
Retinoblastoma (RB) High Retinoblastoma (trilateral disease = co-occurrence of pineoblastoma)
Bone or soft tissue sarcomas
Melanoma
Others
Ophthalmologic examination and imaging; care by a multidisciplinary team
Examination schedule for carriers:
Birth to 8 weeks: Non-sedated examination every 2 weeks
8 weeks to 12 mo: EUA monthly
12–24 mo: EUA every 2 mo
24–36 mo: EUA every 3 mo
36–48 mo: EUA every 4 mo
48–60 mo: EUA every 6 mo
5–7 yr: Examination under sedation every 6 mo
Surveillance for trilateral RB
Brain MRI at the time of RB diagnosis; some centers recommend brain MRI every 6 mo until 5 yr old
Surveillance for second primary tumors
Education regarding second primary tumor risks, and close attention to any new signs or symptoms
Skin examination by a pediatrician during well-child visits, to be continued annually by the primary care physician or dermatologist for melanoma from age 18
Some consider annual WBMR annually after age 8, but no consensus
Radiation avoidance
Knudson's two-hit hypothesis
Usually bilateral or multifocal disease
Mosaic forms, therefore analysis of both tumor tissue and blood is important
RNA studies may also be recommended
Low-penetrance mutations identified
Parent-of-origin effect in some families
RET
Oncogene
Dominant inheritance
Multiple endocrine neoplasia type 2
Three subtypes:

  • MEN2A (Sipple syndrome)

  • Familial medullary thyroid carcinoma (FMTC)

  • MEN2B

High Medullary thyroid carcinoma (MTC) (multifocal, bilateral)
and
primary C-cell hyperplasia
PCC (bilateral) (MEN2A and MEN2B)
Parathyroid adenoma or hyperplasia (hyperparathyroidism, hypercalcemia, renal stones) (MEN2A)
Prophylactic thyroidectomy (varies from infancy to 5 yr of age or older depending on specific RET pathogenic mutation; genotype-phenotype correlations)
Surgical management of identified disease
Serum calcitonin concentration
Consideration of eligibility for treatment with kinase inhibitors
Biochemical screening and imaging for pheochromocytoma and parathyroid abnormalities
Genotype-phenotype correlations
MEN2A: MTC in early-adulthood
FMTC: MTC in middle age
MEN2B: MTC in early-childhood mucosal neuromas (lips, tongue), dysmorphic features (large lips, “marfanoid” body habitus), gastrointestinal ganglioneuromatosis
RET gene also associated with Hirschsprung disease
SDHB, SDHC, SDHD, SDHA, SDHAF2
Tumor suppressors
Dominant inheritance
Hereditary paraganglioma- PCC syndrome High (particularly SDHB and SDHD ) Paraganglioma (PGL)—sympathetic and parasympathetic
PCC
GISTs
Renal clear cell carcinoma
Papillary thyroid carcinoma (unclear data)
Biochemical screening, physical examination, and imaging
Blood pressure at all medical visits starting at age 6–8
Plasma methoxytyramine starting at age 6–8 yr then annually
Plasma free or 24-hour fractionated metanephrines starting at age 6–8 yr then annually
Optional serum chromogranin starting at 6–8 yr then annually
Whole-body MRI skull base to pelvis at 6–8 yr then biennially
Optional: Neck MRI with or without contrast agent starting at age 6–8 then biennially
CBC with RBC indices start at age 6-8 then annual
Secretory and nonsecretory PGL
High blood pressure, headache, anxiety, profuse sweating, palpitations, pallor
SDHB: High risk of malignant transformation
SDHD: Parent-of-origin effect—disease causing when paternally inherited
SDHAF2: Possible parent-of-origin effect (paternal inheritance)
SMARCB1, SMARCA4
Epigenetic regulator
Dominant inheritance
Rhabdoid tumor predisposition syndrome High Rhabdoid tumor
Small cell carcinoma of the ovary hypercalcemic type (SCCOHT)
Schwannomatosis
Brain MRI every 3 mo to age 5
Abdomen: Consider whole-body MRI to age 5, ultrasound every 3 mo
Ovarian and fallopian tumor removal to be considered in carriers before puberty
STK11 (LKB1)
Tumor suppressor
Dominant inheritance
Peutz-Jeghers syndrome (PJS) High Gastrointestinal lesions (PJS-type hamartomatous polyps; colorectal, stomach, small bowel cancer)
Breast
Pancreas
Gynecologic (ovarian, cervix, uterus)
Testes (Sertoli cell)
Lung
Colonoscopy
Upper endoscopy with small bowel visualization
Breast screening with annual mammogram and breast MRI
Investigational-based pancreatic screening
Gynecologic examination
Testicular examination
Mucocutaneous hyperpigmentation of mouth, lips, nose, eyes, genitalia, fingers; may fade after puberty
Females: Sex cord tumors with annular tubules (SCTAT); adenoma malignum of the cervix
TP53
Tumor suppressor
Dominant inheritance
Li-Fraumeni syndrome (LFS) High Soft tissue and bone sarcomas
Breast
Brain
Adrenocortical carcinoma
Leukemia
Others: Gastrointestinal, genitourinary, lung, lymphomas, thyroid, neuroblastoma, skin
Breast screening with annual breast MRI and mammogram beginning in the 20s and 30s, respectively
Optional risk-reducing mastectomy
Physical examination including dermatologic and neurologic examinations
Colonoscopy
Investigational-based whole-body MRI
Consideration of radiation avoidance, if clinically appropriate
Classic LFS criteria
Chompret criteria
Consider genetic testing in BRCA-negative isolated early-onset breast cancer (age <30)
Breast cancers more likely to be estrogen, progesterone, HER2/neu receptor positive
TSC1 and TSC2
Tumor suppressor
Dominant inheritance
Tuberous sclerosis complex (TSC) High Kidney:
Renal cell carcinomas, angiomyolipomas (benign and malignant),
Epithelial cysts,
oncocytoma (benign adenomatous hamartoma)
TSC specialist for management
Imaging
Dermatologic, dental, ophthalmologic examinations; EEG, echocardiogram, ECG
Consideration of eligibility for treatment with mTOR inhibitors
Surgical management
Skin (hypomelanotic macules, facial angiofibromas, shagreen patches, cephalic plaques, ungual fibromas)
Brain (cortical dysplasias, subependymal nodules and subependymal giant cell astrocytomas, seizures, intellectual disability or developmental delay, autism spectrum disorder, psychiatric illness)
Heart (rhabdomyomas, arrhythmias)
Lungs (lymphangioleiomyomatosis)
Genotype-phenotype correlations: Higher risk of renal cell carcinoma in TSC2
Possible association with NETs
VHL
Tumor suppressor
Dominant inheritance
von Hippel-Lindau disease High Clear cell renal cell carcinoma
PCC, paragangliomas
Pancreatic NETs
Eye examination including retina starting at birth then annually
Blood pressure at all medical visits starting at age 2
Plasma free metanephrines or 24-hour urine fractionated metanephrines starting at age 2 then annually
Audiogram starting at age 5 then biennially
MRI of brain with and without contrast agent and MRI of spine with contrast starting at age 8 then biennially
MRI of abdomen starting at age 10 then annually (for RCC and pancreatic NET screen)
Hemangioblastomas of the brain, spinal cord, and retina; renal cysts; pancreatic cysts; endolymphatic sac tumors; epididymal and broad ligament cysts
Genotype-phenotype correlations: VHL type 1, type 2A, type 2B, type 2C with different risks of PCC or renal cell carcinoma

Leukemia and lymphoma predisposition syndromes are not listed. Please refer to dedicated chapters.
AAP, American Academy of Pediatrics; ALL, acute lymphoblastic leukemia; AML, acute myeloblastic leukemia; BSO, bilateral salpingo-oophorectomy; CBC, complete blood count; CNS, central nervous system; ECG, electroencephalogram; EEG, electroencephalogram; EUA, examination under anesthesia; FDG-PET, fluorodeoxyglucose–positron emission tomography; GI, gastrointestinal; GIST, gastrointestinal stromal tumor; HPV, human papillomavirus; HSCT, hematopoietic stem cell transplantation; JMML, juvenile myelo monocytic leukemia; MDS, myelodysplasia; MPNST, malignant peripheral nerve sheath tumors; MRI, magnetic resonance imaging; NET, neuroendocrine tumor; PARP, poly (ADP-ribose) polymerase; PCC, pheochromocytoma; PSA, prostate-specific antigen; RCC, renal cell carcinoma; TSH, thyroid-stimulating hormone; VCE, video capsule endoscopy; WBMR, whole body MRI.

a Earlier initiation of breast cancer surveillance may be warranted in the presence of a significant family history of breast cancer.

b Earlier consideration of risk-reducing salpingo-oophorectomy may be warranted in the presence of a clear family history of ovarian cancer (>1 case).

Box 13.1
Modified from Offit K. Clinical Cancer Genetics: Risk Counseling and Management. New York: Wiley-Liss; 1998, with permission.
Elements of Informed Consent for Genetic Cancer Testing

  • 1.

    What the test is intended to do —that is, determine whether a mutation can be detected in a specific cancer susceptibility gene

  • 2.

    What can be learned from both a positive and negative test result, including information on the magnitude of health risks associated with a positive test result, as well as the risks that may remain even after a negative test result

  • 3.

    The possibility that no additional risk information will be obtained after testing or that the test will result in a finding of unknown significance (e.g., a polymorphism) that might require further studies

  • 4.

    The options for approximation of risk without genetic testing —for example, using empiric risk tables for breast cancer given differing family histories

  • 5.

    The risk of passing a mutation on to children, including options for assisted reproduction (e.g., preimplantation genetics, if discussion is appropriate)

  • 6.

    The importance of notification of family members that they might share a hereditary risk for cancer, with every effort made to assist in contacting family members and providing them access to counseling and testing

  • 7.

    The medical options and limited proof of efficacy for surveillance and cancer prevention for individuals with a positive test result, as well as the accepted recommendations for cancer screening even if genetic testing results are negative

  • 8.

    The technical accuracy of the test —that is, the sensitivity and specificity of the analytic methodology

  • 9.

    The risks of psychologic distress and family disruption whether a mutation is found or not found

  • 10.

    The risk of employment and/or insurance discrimination after disclosure of genetic test results and the level of confidentiality of results compared with other medical tests and procedures

  • 11.

    The risks that nonrelatedness of family members will be discovered and how this information will be disclosed (or not disclosed)

  • 12.

    The fees and costs of testing, including the laboratory test and the associated consultation with the health professional who is providing pretest education, results disclosure, and follow-up, and the costs of preventive procedures, which might not be covered by third-party payers

Mutations (i.e., pathogenic variants) in genes whose alterations result in susceptibility to cancer may be inherited, whereby a family history usually includes multiple individuals diagnosed with cancer, or may occur de novo in families in which the history of cancer may be unremarkable. In many instances, the same genes recurrently mutated in specific sporadic cancers, if mutated in the germline, are also susceptibility genes for the same types of tumors. This phenomenon reflects the genotype-phenotype relationships that arise as a result of aberration of particular biologic pathways and thus define the patterns that characterize hereditary cancer syndromes. Increasingly, specific pathologic hallmarks have been associated with cancer predisposition syndromes (see Table 13.1 ). It is now widely accepted that cancer pathogenesis derives from inherited or acquired alterations in a set of “driver” genes that are associated with abnormal cellular function, resulting in uncontrolled cell division and/or loss of the normal fidelity of DNA repair and replication. This process results in accumulation of further mutations characterized as “passengers” in the neoplastic process which contribute to tumor survival. With the advent of massively parallel sequencing, progress in the understanding of tumor genetics has catalogued these “driver” and “passenger” mutations, illuminating biologic pathways and facilitating development of therapies targeted toward these pathways. At the same time, the definition of pathways perturbed in malignant transformation and progression has provided novel biomarkers of utility for disease prognosis and, as is the focus here, disease susceptibility.

The syndromes of cancer susceptibility included in this chapter are those that are most commonly encountered in oncologic practice, as well as several recently defined entities, some of which have been associated with targeted therapies. The most common of these syndromes, predisposing to cancers of the breast, ovary, colon, prostate, and pancreas, affect tens of thousands of Americans who are diagnosed with these cancers each year in the United States and result in an increased risk for a second neoplasm for millions of cancer survivors.

Inherited cancer predisposition can be considered as a spectrum, arising from single or combined low-, moderate-, and high-risk genetic variants for which the timing of disease onset is likely modified by the type of genetic variant and its effect on normal cellular function and the responses to environmental factors. In general, highly penetrant hereditary cancer syndromes account for about 5% to 10% of most types of cancer and are caused by rare genetic variants. We are now also aware of dozens of “moderate-penetrance” genes that in general confer a modest degree of cancer risk, with a relative risk (RR) ranging between 2 and 5 ( Table 13.3 ). Although screening for such moderate-penetrance genes was previously not routinely undertaken, with the availability of next-generation sequencing, screening for mutations in many genes simultaneously, often in multigene panels, has become readily available. The value of screening for moderate-penetrance genes remains controversial because neither the clinical validity, the accuracy with which a genetic test predicts the development of cancer, nor the clinical utility, the degree to which the use of the genetic test informs clinical decision making and leads to improved health outcomes, has been clearly proven. In addition to the high- and moderate-penetrance cancer genes, for nearly all the common malignancies, hundreds of additional genetic loci have been identified, largely via genome-wide association studies, with each genetic variant, usually single-nucleotide polymorphisms (SNPs), being associated only a modest increased risk of cancer (RR, ~1.1–1.5). Given the limited clinical validity and utility of SNPs with such small effect sizes, clinical testing for individual risk loci is currently performed, although research efforts to devise polygenic risk models are underway and have already been incorporated into models of risk stratification for BRCA mutation–carrying males and females. In addition, these frequent but low-penetrance variants identified through genome-wide association studies are thought to explain a portion of the excess familiality of cancer. To account for the missing familiarity that remains, sequencing of protein coding regions of the genome or the entire genome itself has led to the discovery of novel cancer susceptibility genes. Together, the knowledge of rare and common variations has the potential to inform the clinician about the combined effect of genetic factors that lead to cancer, extending beyond risks conferred by single genes. Ultimately, to assess a person's risk of cancer, one will need to combine genetic factors within the context of environmental factors, which also may include exposures to cancer therapies.

Table 13.3
Selected Moderate-Penetrance Genes and Associated Malignancies
Data from Tung N, Domchek SM, Stadler Z, et al. Counselling framework for moderate-penetrance cancer-susceptibility mutations. Nat Rev Clin Oncol. 2016;13(9):581-8; Ma X, Zhang B, Zheng W. Genetic variants associated with colorectal cancer risk: comprehensive research synopsis, meta-analysis, and epidemiological evidence. Gut. 2014;63(2):326-36; and Liang J, Lin C, HuF, et al. APC polymorphisms and the risk of colorectal neoplasia: a HuGE review and meta-analysis. Am J Epidemiol . 2013;177(11):1169-79.
Breast Ovary Colorectal Pancreas Melanoma Prostate Renal Other
APC*I1307K X
RR, 2.17
(95% CI, 1.6–2.9) a
ATM X
RR, 2.8 (90% CI, 2.2–3.7)
X Autosomal recessive
Ataxia telangiectasia
BARD1 (ID) (ID)
BRIP1 X
RR 11.2 (95% CI, 3.2–34.1)
in case control;
RR, 3.41 (95% CI, 2.1–5.5) in segregation analysis
Autosomal recessive
Fanconi anemia
CHEK2 X
Truncating:
RR, 3.0 (90% CI 2.6–3.5);
Missense (I157T):
RR, 1.58 (95% CI, 1.4–1.8)
(ID) X
1100delC:
RR, 1.88 (95% CI, 1.3–2.7);
I157T:
RR, 1.56 (95% CI, 1.3–1.8)
(X)
MAX X (Pheochromocytoma, paraganglioma)
MITF X (X)
MRE11A (ID) (ID)
MUTYH (monoallelic) X
RR, 1.17 (95% CI, 1.0–1.3)
NBN (X)
c.657del5:
RR, 2.7 (90% CI 1.9–3.7)
Autosomal recessive Nijmegen breakage syndrome
NF1 X High penetrance Neurofibromatosis type 1
RAD50 (ID) (ID)
RAD51C, RAD51D X
RAD51C: RR, 5.2 (95% CI, 1.1–24); RAD51D: RR, 12 (95% CI, 1.5–90)
RAD51C:
Autosomal recessive Fanconi Anemia
TMEM127 X (Pheochromocytoma)
Risk estimates listed assume no family history of the implicated cancer.
(X): To date, limited evidence to support the association and the strength of the association is unclear.
(ID): Insufficient data. Occurrence documented, but association not clearly supported by evidence.
CI, Confidence interval; RR, average relative risk.

a Risk estimate for APC*I1307K is based on Ashkenazi Jewish population.

Common Syndromes of Cancer Predisposition

The major syndromes of cancer predisposition that affect adults include breast, ovarian, colon, prostate, gastric, and pancreatic cancer, as well as a number of other less common but equally important cancer predispositions. Some of these syndromes, including multiple endocrine neoplasia type 1 (MEN1), retinoblastoma, melanoma, von Hippel-Lindau (VHL) disease, and familial pheochromocytoma and paraganglioma, are described elsewhere in this text. Comprehensive reviews of these cancer predispositions are offered elsewhere, including Li-Fraumeni syndrome, MEN1, retinoblastoma, melanoma, and VHL disease. Some syndromes such as neurofibromatosis also have cancer predispositions. A detailed outline of the elements of setting up a comprehensive cancer genetic and genomics counseling service has also been published.

Breast and Ovarian Cancer Syndromes

Clinical Features

Although only about 18,000 cases of breast cancer each year are associated with an obvious hereditary predisposition, primary cancers developed in more than 200,000 breast cancer survivors in the United States as a result of a hereditary predisposition, and these survivors remain at risk for secondary cancers. Genetic testing has emerged as one of the most important indicators of risk factors, pointing to a need for intensified screening for breast cancer. When detected at an early stage, more than 90% of breast cancers are curable. These statistics underscore the rationale for the use of genetics in clinical oncology. We have previously reviewed in detail the management of women and men who are at hereditary risk for breast cancer, and this review is summarized and updated here.

From 1 in 150 to 1 in 800 individuals in the population carry a genetic susceptibility to breast cancer, and the prevalence is much higher in certain ethnic groups. Syndromes of breast cancer susceptibility are linked to mutations of BRCA1 and BRCA2, in addition to a smaller number of cases with germline mutations of ATM, PALB2, PTEN, p53, CHEK2 , STK11 , CDH1, and rarer syndromes ( Tables 13.4 and 13.5 ). Cowden syndrome (CS) was initially described as a dominant inheritance of multiple hamartomatous lesions, including papillomas of the lips and mucous membranes and acral keratoses of the skin. This disease was ultimately linked to germline mutations of PTEN and is discussed separately later. In persons with Li-Fraumeni syndrome, early-onset breast cancer occurs with soft tissue sarcomas, osteosarcoma, leukemia (particularly hypodiploid acute lymphoblastic leukemia), brain tumors, adrenal cortical tumors, and other cancers. Rarely, atypical breast-ovarian kindred may be found to have a germline p53 mutation. Although rare, women with Peutz-Jeghers syndrome, associated with germline mutations in the STK11 gene, are at increased risk for breast cancer as well as a variety of other malignancies. More recently, mutations in PALB2, partner and localizer of BRCA2, have been implicated in both familial breast and pancreas cancer risk.

Table 13.4
Genes Associated With Hereditary Breast Cancer (BC) Predisposition
Gene Syndrome Relative Risk of BC BC Risk by Age 80 Yr Associated Cancers
HIGH PENETRANCE
BRCA1 HBOC ~15–30 70% Ovarian, other
BRCA2 HBOC ~10–20 70% Ovarian, pancreatic, prostate, other
p53 Li-Fraumeni syndrome 100 50% by 60 yr Soft tissue sarcoma, osteosarcoma, brain tumors, adrenocortical carcinoma, leukemia, other
PTEN Cowden syndrome No reliable estimate 70%–80% Thyroid (follicular and rarely papillary) endometrial, genitourinary, other
Bannayan-Riley-Ruvalcaba syndrome
Proteus
Proteus-like syndrome
STK11 Peutz-Jeghers syndrome No reliable estimate 30% by age 60 Small intestine, colorectal, uterine, testicular and ovarian sex chord tumors, other
CDH1 Hereditary diffuse gastric carcinoma ~3.25 39% Lobular breast, diffuse gastric, other
LOWER OR MODERATE PENETRANCE
ATM (heterozygote) Ataxia-telangiectasia in homozygotes ~3 30 Undefined in heterozygotes
CHK2 (CHEK2) Li-Fraumeni variant 1.5-3 20%–30% Undefined
PALB2 None known 5 ~40 Undefined in heterozygotes
HBOC, Hereditary breast and ovarian cancer syndrome.

Table 13.5
Genes Associated With Hereditary Ovarian Cancer (OC) Predisposition
Gene Syndrome Relative Risk of OC OC Risk by Age 80 Yr Associated Cancers
HIGH PENETRANCE
BRCA1 Hereditary breast ovarian cancer syndrome ~50 ~40% Breast, other
BRCA2 Hereditary breast ovarian cancer syndrome ~8 11%–26% Breast, pancreas, prostate, other
MLH1
MSH2
MSH6
PMS2
EPCAM
Lynch syndrome ~4 ~20% Colon
Uterine
Stomach
Small intestine
Urinary tract
Pancreatic
Possible other sites
LOWER OR MODERATE PENETRANCE
RAD51C None ~5 ~6% Undefined
Autosomal recessive Fanconi anemia
RAD51D None ~12 ~14% Undefined
Autosomal recessive Fanconi anemia

Early linkage studies suggested that in about 50% of breast cancer kindreds, the cancer was linked to BRCA1 ; in 30%, it was linked to BRCA2 ; and in the remainder, it was linked to other either identified or unidentified genes. In up to two-thirds of families with both male and female breast cancer, the cancers were due to BRCA2, whereas more than 80% of families with both breast and ovarian cancer harbored mutations in BRCA1. BRCA1 -linked breast cancers are associated with a basal-like subtype and higher mitotic indices. They are more likely to be of higher grade, are more frequently estrogen and progesterone receptor negative, and demonstrate a basal-like phenotype. They display a more “aggressive” phenotype, including a higher proportion of cells in S phase, and other indices.

Multiple moderate-penetrance breast cancer susceptibility genes have also been identified (see Table 13.4 ). In Northern European families, specific mutations of CHEK2 are associated with familial breast cancer, with pathogenic CHEK2 variants being present in ~1% of Caucasians of European descent. Although the common European mutation in CHEK2 is rare in persons in North America, analysis of women for CHEK2 founder mutations stratified for family history of breast cancer demonstrated that carriers with a positive family history had a greater than 25% lifetime risk for breast cancer. Important to note, degree of breast cancer risk in women with mutations in CHEK2 appears to be dependent on both the nature of the specific mutation (truncating versus missense) and the strength of the family breast cancer history. Carriers of mutations in ATM, another moderate-penetrance breast cancer susceptibility gene, have an approximately twofold elevated breast cancer risk.

Multiple, common, lower-penetrance genes are likely to account for a significant component of currently unexplained familial breast cancer risk. A host of putative lower-penetrance gene mutations have been identified through whole-genome association studies, although the clinical relevance of these associations remains unclear. In a large cohort of breast cancer patients referred for genetic testing, approximately 10% had a related germline mutation.

Compared with the 10% breast cancer risk for women in the general population, estimates of the breast cancer risk that is conferred by a common high-penetrance susceptibility gene ranged from 67% to 69% by age 70 years based on epidemiologic analyses. This was confirmed in a prospective study that defined breast cancer risk by age 80 of 72% and 69% for BRCA1 and BRCA2 carriers, respectively. This study also reported a contralateral breast cancer risk, up to 20 years after diagnosis, of 40% and 26% for BRCA1 and BRCA2, respectively. In the setting of a mutation in PALB2, the breast cancer risk to age 80 is approximately 45%, whereas for those with mutations in ATM and CHEK2 it is less than 30%.

Epithelial ovarian carcinoma can be delineated into six distinct histologic subtypes—high-grade serous, carcinosarcoma, clear cell, endometrioid (International Federation of Gynecology and Obstetrics [FIGO] grades 1–3), mucinous, and low-grade serous—with differing manifestations, prognoses, and molecular and immunohistochemistry profiles. The high-grade serous subtype comprises the majority of cases of ovarian cancer. Mutations in BRCA1 and BRCA2 show a strong association with the high-grade serous carcinoma of ovary and fallopian tube or peritoneal origin, such that germline testing for BRCA1 and BRCA2 together can have up to a 25% detection rate in this histologic subtype. Germline testing for BRCA1 and BRCA2 is standard of care for all patients with a high-grade serous ovarian cancer diagnosis. Mutations in the mismatch repair (MMR) genes are also associated with ovarian cancer risk, usually endometrioid histology; this is further discussed in the later section on Lynch syndrome (LS ). Similar to breast cancer, a number of more moderate-penetrance ovarian cancer susceptibility genes have also been identified. BRIP1, RAD51C, and RAD51D are associated with approximately a 5-fold to 12-fold risk of ovarian cancer (see Table 13.5 ). Earlier data suggested an association with BRIP1 and breast cancer, but a large population-based study demonstrated that truncating mutations in this gene were not associated with a substantial increase in breast cancer risk.

In a study of close to 2000 women with ovarian cancer, 18% of patients had a germline mutation in a gene associated with ovarian cancer risk. The majority of patients in this study had high-grade serous histology (78%), but all histologic subtypes were represented. As expected, mutations in BRCA1 and BRCA2 accounted for the bulk of inherited risk, with mutations in BRIP1, RAD51C, RAD51D, and the MMR genes accounting for 3% cumulatively. Compared with a population risk for ovarian cancer of approximately 1.5%, mutations in BRCA1 and BRCA2 are associated with a risk of 44% and 17% by age 80, respectively. For BRIP1, RAD51C, and RAD51D, the cumulative lifetime risk by age 80 is 4% to 14%.

The role of ascertainment and other possible biases in deriving cancer risk estimates, as well as risks for cancer of the prostate, colon, pancreas, and other sites in BRCA mutation carriers has been reviewed. In addition to Fanconi anemia, persons with compound BRCA2 mutations may experience childhood medulloblastomas. Heterozygous mutations in BRCA1 and BRCA2 harbored by pediatric cancer patients have been detected through tumor normal sequencing; however, it is not clear at this point if these are cancers associated with the familial syndrome or simply true and unrelated, and this is an active area of research.

Genetics

BRCA1 and BRCA2 are both important for DNA repair, in particular homologous recombination, which enables repair of double-stranded DNA breaks. BRCA1 is a large gene, spanning more than 100,000 bases of genomic DNA with 22 coding and 2 noncoding exons. BRCA2 is also large, consisting of 27 exons across 70 kb of genomic DNA. Both genes, by coincidence, have a large exon 11. An update of reported mutations is accessible through the Internet at http://research.nhgri.nih.gov/bic/ .

Most BRCA1 mutations cause premature truncation of the peptide by frameshift or nonsense sequence changes. Large germline rearrangements also occur in BRCA1 and BRCA2. Five percent to 10% of BRCA mutations that are missense are problematic because they are of unknown clinical significance. The proportion of these variants that are of unknown significance was as high as 10% to 23% in some series, posing counseling challenges. The role of BRCA1 and BRCA2 in DNA damage response and homologous recombination is reviewed elsewhere.

Founder BRCA mutations have been documented in genetically isolated populations. In North American families, the most common founder mutations occur in persons of Ashkenazi Jewish origin. These mutations include a two-base-pair deletion in codon 23 of BRCA1, termed 185delAG (c.68_69delAG); another mutation in BRCA1, 5382insC (c.5266dupC); and the 6174delT (c.5946delT) mutation in BRCA2. About 1 in 40 Ashkenazi Jews harbor one of the common BRCA1 or BRCA2 mutations, a relatively high carrier frequency for an inherited cancer predisposition syndrome. Other mutations in BRCA1 and BRCA2 occur in the Ashkenazim; 16 of 737 Ashkenazi Jews who were tested in a clinic-based ascertainment had a nonfounder mutation (2%). In another study of Ashkenazi individuals with a personal history of breast or ovarian cancer who had previously been shown not to have a founder mutation, 3 of 70 (4.3%) had a deleterious nonfounder mutation. This very low nonfounder BRCA rate in Ashkenazim was also confirmed again in a 2017 study. Founder mutations in populations other than the Ashkenazim have also been observed.

Other Genes

Protein truncating variants in PALB2, CHEK2, and ATM have been associated with increased breast cancer risk as outlined earlier. The risk associated varies per study and variant. A Finnish founder mutation in PALB2 c.1592delT is associated with higher than population risk of breast cancer, but risk estimates are lower than those reported with other PALB2 mutations. Most of the data for CHEK2 and increased breast cancer risk relates to the truncating c.1100delC variant found in Northern Europeans. The missense variant CHEK2 I157T is associated with a very modest increased risk (RR, 1.58; 95% confidence interval [CI], 1.42–1.75), with some laboratories not reporting this variant as being pathogenic. The autosomal recessive neurologic disorder ataxia telangiectasia is associated with compound heterozygote mutations in ATM, a gene involved with DNA double-strand break repair and cell cycle control . Although this is a rare event, monoallelic ATM mutations are relatively common, occurring in approximately 3% of Caucasians. Women with heterozygous ATM mutations have a moderate, approximately double population risk of breast cancer as noted earlier. BRIP1, BRCA1 interacting protein C-terminal helicase 1 gene, is part of the DNA repair machinery. A founder mutation, BRIP1, c.2040_2041insTT was associated with increased risk of ovarian cancer in the Icelandic population. Subsequently, case-control series reported increased ovarian cancer risk for other truncating BRIP1 mutations. RAD51C and RAD51D are part of the Fanconi- BRCA homologous recombination repair pathway and have been associated with a modest increased risk of ovarian cancer.

Clinical Management

Three elements of breast surveillance that are recommended to women with increased risk of breast cancer are self-examination, clinician examination, and imaging. The evidence base underlying these recommendations has been reviewed, and updated guidelines are available.

Increasingly, breast cancer screening including mammography and magnetic resonance imaging (MRI) together has been shown to have greatest sensitivity. For women who are at the highest hereditary risk for breast cancer, whose breasts are difficult to examine, or who have had biopsy results showing atypia, it might be appropriate to discuss the option of removing the healthy breasts as a preventive measure (i.e., prophylactic mastectomy). Retrospective studies and reviews document the well-established risk-reducing role of surgery in high-risk patients. Prospective and combined consortium studies have also shown the efficacy of this approach.

Because of their antiestrogen properties, tamoxifen, raloxifene, and aromatase inhibitors have been shown to decrease breast cancer rates in persons who are at increased risk. Two studies on the impact of tamoxifen on subsequent breast cancer risk in BRCA1 and BRCA2 mutation carriers have shown conflicting conclusions, although only one of these studies was sufficiently powered to reach significant results. Tamoxifen was confirmed to decrease contralateral breast cancer risk in a follow-up of one of these studies of BRCA mutation carriers. Providers are encouraged to discuss breast cancer chemoprevention with women who are at increased risk of breast cancer. There are no studies demonstrating a mortality benefit for this approach, and a careful discussion of potential benefit versus toxicity is required.

Small trials have demonstrated the ability of ultrasound with Doppler and CA125 to reveal early-stage ovarian cancers in BRCA mutation carriers. However, the efficacy of this approach has not been proven in large, prospective studies, and the US Food and Drug Administration (FDA) has recommended against screening for ovarian cancer in all women, including those at high risk. Therefore prophylactic removal of the ovaries and fallopian tubes is recommended to women with strong family histories of ovarian cancer, in families linked to BRCA1 or BRCA2, or to women who are considering hysterectomy in the setting of a germline mutation associated with LS. Studies examining specimens from prophylactic salpingo-oophorectomy in BRCA1 and BRCA2 mutation carriers have implicated the fimbriated part of the fallopian tube as harboring precursor lesions to high-grade serous epithelial ovarian cancer. Thus prophylactic salpingo-oophorectomy is recommended because such surgeries not only decrease the incidence of subsequent breast and ovarian cancer but also reveal occult early-stage pelvic neoplasms, decreasing mortality. These studies have also confirmed that high-grade serous carcinoma can still occur after prophylactic oophorectomy, albeit at much reduced rates. For women with high-penetrance risk genes, risk-reducing prophylactic oophorectomy is recommended when childbearing is complete or at approximately age 40 . Given the lack of an adequate screening modality, difficulty in diagnosis, and resultant late stage at presentation and the associated mortality with invasive ovarian cancer, risk-reducing surgery is also recommended for carriers of the more modest-risk mutations in genes BRIP1, RAD51C, and RAD51D. However, given the later age of onset of ovarian cancer compared with BRCA1 and BRCA2, risk-reducing surgery may be delayed until patients in this group are perimenopausal or postmenopausal.

Combination oral contraceptives that contain estrogen and high-dose progestin result in a time-dependent, protective effect against ovarian cancer in some but not all studies of BRCA mutation carriers. There remains the concern of a small increased risk of breast cancer due to oral contraceptives in this group, particularly in persons with BRCA1 mutations.

Germline BRCA1 and BRCA2 mutation–associated tumors have defective DNA repair mechanisms because of the loss of function of the remaining normal copy of the gene. This inability of the tumor to repair its DNA effectively can be exploited by therapies that disrupt alternative DNA repair pathways, such that the tumor cell accumulates so much DNA damage that it can no longer survive. Inhibitors of the base excision repair enzyme poly (ADP-ribose) polymerase (PARP inhibitors) use this principle of synthetic lethality and are FDA approved for women with recurrent ovarian cancer in a number of clinical settings. Although benefit has been demonstrated in all patients with ovarian cancer, it is substantially greater in patients with a germline or somatic BRCA1/2 mutation. A recent randomized trial has also demonstrated that patients with metastatic breast cancer and germline BRCA1/2 mutations had significant progression-free survival benefit with olaparib.

Cowden Syndrome

Clinical Features

CS is an autosomal dominant disorder characterized by multiple hamartomas with a high risk of benign and malignant tumors of the thyroid, breast, and endometrium. Consensus criteria for CS establish three diagnostic categories.

Pathognomonic criteria include adult Lhermitte-Duclos disease (LDD) (defined as presence of a cerebellar dysplastic gangliocytoma) and mucocutaneous lesions, facial trichilemmomas, acral keratoses, papillomatous lesions, and mucosal lesions. Major criteria include breast cancer, thyroid cancer (especially follicular histology), macrocephaly, and endometrial carcinoma. Minor criteria include other thyroid lesions (e.g., goiter), mental retardation, hamartomatous intestinal polyps, fibrocystic breast disease, lipomas, fibromas, and genitourinary tumors (e.g., uterine fibroids and renal cell carcinoma) or genitourinary malformation. The operational diagnosis of CS is made if an individual meets any one of the following criteria: pathognomonic mucocutaneous lesions alone if there are six or more facial papules, of which three or more must be trichilemmoma, or cutaneous facial papules and oral mucosal papillomatosis; or oral mucosal papillomatosis and acral keratoses; or six or more palmoplantar keratoses. Alternatively, the individual may fulfill two major criteria, but one must include either macrocephaly or LDD. Alternatively, the individual may have one major and three minor criteria or four minor criteria.

The palmar and plantar hyperkeratotic pits usually become evident later in childhood. Subcutaneous lipomas and cutaneous hemangiomas are seen in persons with CS with low frequency. An increased risk of early-onset male breast cancer has been noted in mutation carriers.

Cumulative lifetime risks for female breast cancer are 81% to 85.2% ; for LDD, 32% (CI, 19%–49%) ; for thyroid cancer, 21%–35.2%, ; for endometrial cancer, 19%–28.2% ; and for renal cancer, 15%–33.6% (CI, 6%–32%). The lifetime risks for colorectal cancer (CRC) are 9% to 16%, and the risk for melanoma is 6%.

Genetics

The CS-linked PTEN was mapped to 10q22-23. Bannayan-Riley-Ruvalcaba syndrome, characterized by macrocephaly, intestinal polyps, lipomas, and pigmented penile macules, is also caused by germline mutations in PTEN .

PTEN acts as a tumor suppressor by mediating cell cycle arrest, apoptosis, or both. Full sequencing and deletion/duplication analysis are available clinically, and promoter analysis is available on a research basis. Heterozygous germline mutations in PTEN cause most cases of CS. Nonsense, missense, and frameshift mutations that are predicted to disrupt normal PTEN function have been identified in some families, including mutations that disrupt the protein tyrosine/dual-specificity phosphatase domain. Initially it was shown that PTEN mutations can be detected in about 80% of patients with CS. More recently, the mutation detection rate in persons meeting CS criteria was found to be 34%, raising the possibility that current aspects of testing criteria are too broad.

Risk Management Recommendations

Persons with known germline PTEN mutations should undergo appropriate cancer screening. In view of recent work outlining cancer risks in persons with CS, some investigators have suggested that National Comprehensive Cancer Network (NCCN) management guidelines be modified to include renal cancer screening using biannual renal imaging from age 40 years or 5 years earlier than the earliest kidney cancer diagnosis in the family, and endometrial sampling in adulthood or 5 years earlier than the earliest endometrial cancer diagnosis in the family. Female patients with CS should have breast self-examination training and education starting at age 18 years. Beginning at age 25 years, clinical breast examinations should be performed every 6 to 12 months, and annual mammography and MRI screening should start at age 30 years or 5 years before the earliest age of breast cancer diagnosis in the family. Men should perform monthly breast self-examination. Female patients should receive endometrial cancer screening beginning around age 30 years or 5 years before the earliest age of endometrial cancer diagnosis in the family. Persons with CS should undergo biannual colonoscopy from age 40 years. Both men and women should have a comprehensive annual physical examination starting at diagnosis with screening for skin and thyroid lesions, including a baseline and annual thyroid ultrasound and dermatologic examination. Finally, preimplantation genetic testing for CS can be performed if a mutation is described in a parent.

Common Colon Cancer Predisposition Syndromes

High-penetrance cancer susceptibility syndromes account for about 5% of CRC cases. The most common syndrome is LS, with familial adenomatous polyposis (FAP) constituting a rarer familial syndrome. Genetic epidemiologic analyses suggest a common susceptibility allele for both colon cancer and adenomatous polyps that accounts for at least 15% and possibly half of cases.

Lynch Syndrome

A constellation of colon and endometrial cancers became known as Lynch syndrome, formerly referred to as hereditary nonpolyposis colon cancer (HNPCC). LS is an autosomal dominant syndrome caused by mutations in one of four MMR genes (MLH1, MSh2, MSH6, PMS2), or a gene located nearby (EPCAM), with resultant errors in DNA replication yielding a microsatellite instability (MSI) phenotype. LS accounts for 3% of all CRCs.

Clinical features

LS-associated CRC is characterized by an accelerated progression of the adenoma-to-carcinoma sequence, with a predominance of right-sided colorectal tumors with pathology demonstrating poorly differentiated adenocarcinoma with mucinous and signet ring cell features, a Crohnlike reaction, and tumor-infiltrating lymphocytes, in addition to presentation with metachronous or synchronous colorectal tumors. Patients with LS have a 50% to 80% lifetime risk of colon cancer, with a median age of diagnosis of 44 years in the proband and 61 years in mutation carrier relatives. A 40% to 60% endometrial cancer risk by age 70 years has been reported, with a median age at onset from the late 40s to early 50s, as compared with a 3% risk for endometrial cancer in the general population. LS also confers a high lifetime risk for ovarian cancer (12%–15%). Five additional tumor sites demonstrated increased observed/expected (O/E) ratios in LS kindreds: cancers of the stomach (O/E = 4.1), small intestine (O/E = 25), ureter (O/E = 22), and kidney (O/E = 3.2). More recently, adrenocortical neoplasm have also been implicated in LS and, although still controversial, some studies also have demonstrated a modest increased risk of prostate and breast cancer in patients with LS.

At a 1991 meeting in Amsterdam, the International Collaborative Group on LS defined the syndrome as (1) histologically verified CRC in three or more relatives, including a first-degree relative of the other two; (2) CRC involving at least two generations; and (3) one or more CRCs diagnosed before age 50 years. The subsequent “Amsterdam II criteria” for LS were redefined to include extracolonic LS-associated cancers. In 1996, the “Bethesda criteria” delineated individuals at risk for LS for whom molecular genetic analysis may be considered. A revised set of Bethesda Guidelines was developed to identify subjects who are at high risk of having a germline MMR gene mutation. Multivariate logistic regression risk models using personal and family medical histories estimate the probability of carrying an LS mutation. In the most common LS-associated cancers, including CRC and endometrial cancers, the need for tumor screening for LS has traditionally been based on the age at cancer diagnosis or on the strength of the personal and family history, as outlined in the Revised Bethesda Guidelines. However, per the most recent guidelines from EGAPP (Evaluation of Genomic Applications in Practice and Prevention Working Group) and NCCN, screening of all colorectal tumors (or at least all colorectal tumors in patients younger than 70 years at diagnosis or those older than 70 years who meet Revised Bethesda Guidelines) and endometrial tumors via MSI or MMR protein immunohistochemical staining analysis is now recommended.

Genetics

About 45% to 70% of LS families harbor mutations in one of the following four genes: MSH2, MLH1, MSH6, and PMS2. In addition, germline deletions of the 3′ region of the EPCAM gene, just upstream of MSH2, also result in LS, through silencing of the MSH2 gene. Mutations of MSH2 and MLH1 were far more frequent than the others, accounting for about 30% each of families meeting Amsterdam criteria for LS. More than 75% of mutations in MSH2 and MLH1 were inactivating insertions, deletions, alterations in premessenger RNA splicing signals, and nonsense mutations. Of 120 mutations surveyed, 23% were missense mutations. Mutations of MSH6 less frequently result in the “replication error repair” phenotype but account for a significant number of familial colon cancer families. The replication error repair phenotype is commonly detected as MSI with use of polymerase chain reaction screening of tumors with microsatellite markers. The MSI phenotype is present in about 80% of LS-associated colon cancers and in about 15% of sporadic colon tumors, as well as in other tumors associated with LS (e.g., uterine and gastric cancers). MSI results in a genome-wide increased mutation rate, causing mutations in oncogenes, tumor suppressors, and microsatellite regions.

For patients with an abnormal screening test result (MSI-high or MMR-deficient tumor), various algorithms have been developed to help guide subsequent evaluations, which may include germline genetic testing or further tumor analyses, such as MLH1 promoter hypermethylation and/or BRAF*V600E somatic mutation in MLH1/PMS2 protein deficient cases. In older patients with CRC, hypermethylation of the MLH1 promoter, or presence of a BRAF*V600E somatic mutation, may account for the lack of MLH1 protein expression. This epigenetic (nonhereditary) mechanism of MLH1 promoter hypermethylation is responsible for most of the remaining patients whose tumors are characterized by defective DNA MMR.

Biallelic mutations in the same MMR genes result in constitutional MMR deficiency, which classically manifests as childhood-onset cancers, in particular hematologic malignancies, brain tumors, and early-onset CRCs and café au lait macules similar to those seen in neurofibromatosis type 1, although a milder phenotype can exist in persons with biallelic mutations in PMS2 as evidenced by reports of first cancer diagnoses in mutation carriers in the third and fourth decades.

Clinical management

For probands diagnosed with LS, observational data have shown that surveillance colonoscopy lowers CRC incidence and mortality by more than 50% and improves patient survival. Persons with LS are advised to undergo colonoscopic surveillance every 1 to 2 years, preferably annually, starting at age 20 to 25 years. Some guidelines suggest that for MSH6 mutation carriers, screening colonoscopy may be delayed to the age 30 years. A baseline upper endoscopy should be performed, but the optimal subsequent screening interval has yet to be established. Prophylactic subtotal colectomy sparing the rectum should be considered after the first CRC diagnosis, given the high rate of metachronous CRCs. No benefit has as yet been demonstrated with other screening strategies directed toward the LS-associated malignancies. Endometrial cancer screening entails endometrial biopsy at age 30 to 35 years. Annual urinalysis with cytologic evaluation or urinalysis for microhematuria may also be considered, especially in MSH2 carriers, in whom urothelial risk is highest, but minimal data support the efficacy of this screening test in early urothelial tumor detection. A retrospective study of prophylactic bilateral oophorectomy and hysterectomy showed protection from uterine and endometrial cancer, although the estimates were influenced by a retrospective study design. Nonetheless, a combination of risk-reducing bilateral oophorectomy and hysterectomy is a reasonable option after childbearing or at the time of colorectal resection for a CRC occurring in women with LS.

Identification of LS at the time of CRC diagnosis can affect clinical management. Subtotal (versus segmental) colectomy can reduce the increased risk of metachronous CRC in persons with LS by 31% for every additional 10 cm of bowel removed. Tumors that demonstrate the MSI-H phenotype may not benefit from 5-fluorouracil–based chemotherapy and have improved stage-independent survival compared with proficient-MMR CRC.

More recently, presence of MMR deficiency, a nearly universal finding in LS-associated tumors, predicted for response to immune checkpoint blockade. Specifically, in patients with advanced MMR-deficient CRC, a response rate of 62% was seen, as compared with none in MMR-proficient tumors, to pembrolizumab, an anti–programmed death 1 immune checkpoint inhibitor. In further evaluation of immune checkpoint blockade across 12 different MMR-deficient malignancies, a comparable high response rate was observed, including response in LS patients. Checkpoint blockade has also been shown to be of benefit for children with biallelic MMR deficiency. Future studies are evaluating the role of immunotherapy in patients with early-stage MMR-deficient tumors.

Polyposis Syndromes

Familial adenomatous polyposis

Clinical features.

FAP (adenomatous polyposis coli) manifests with hundreds to thousands of adenomatous polyps at a young age. The hallmark of FAP is the development of more than 100 adenomatous colonic polyps in the teenage years with a resultant risk of CRC of approximately 90% by age 45 years. Colon cancer is therefore inevitable if the colon is not removed; cancer occurs at an average age of 39 years. Flexible sigmoidoscopy at an early age establishes the diagnosis, and prophylactic colectomy is performed in the teen years. Patients remain at risk for primary adenomas and carcinomas of the duodenum and rectum, as well as thyroid carcinoma, hepatoblastoma, and other hepatopancreatic tumors. Gastric fundic gland polyps occur and are often dysplastic, although an increased risk of gastric cancer has yet to be definitely demonstrated in Western populations. Duodenal polyps are also prevalent, with a lifetime risk of duodenal cancer, particularly ampullary cancer, of 5% to 12%. Other FAP-associated cancers are listed in Table 13.2 . Benign extraintestinal manifestations of FAP include desmoid tumors, osteomas of the jaw and dental anomalies, congenital hypertrophy of the retinal pigment epithelium (CHRPE), lipomas, fibromas, sebaceous and epidermoid cysts, and nasopharyngeal angiofibromas.

Genetics.

FAP is an autosomal dominant syndrome that occurs in 1 of every 7000 to 38,000 individuals and arises from germline mutations in the APC gene. Twenty-five percent of mutations are de novo, whereas a small fraction of cases result from somatic mosaicism. Genotype-phenotype correlations have been described, particularly for CHRPE and the number of polyps observed. An attenuated form of FAP is associated with mutations at the extreme 5′ and 3′ ends of the gene. Up to 30% of patients with multiple adenomas (15–100 adenomas) who test negative for APC mutations carry biallelic mutations in MUTYH , discussed in detail later in this chapter.

Clinical management.

FAP may manifest as childhood malignancy, and thus the diagnosis is important as early as possible in a proband's lifetime, even prenatally via preimplantation genetic testing if in vitro fertilization was used. Children are at risk for hepatoblastoma for the first 5 years of life and are screened accordingly. Annual flexible sigmoidoscopy for CRC screening begins at age 10 to 12 years. Once polyps are identified, annual colonoscopic surveillance is recommended. Prophylactic proctocolectomy, usually in the late teens, is the treatment of choice. Sulindac, a nonsteroidal antiinflammatory drug, and selective cyclooxygenase-2 inhibitors reduce adenoma size and number and may be considered for chemoprevention to delay but not preclude surgical intervention. Patients require lifelong surveillance for extracolonic tumors, including tumors of the upper gastrointestinal tract and the ileal pouch (if proctocolectomy is performed). The burden and histologic features of duodenal polyps determines the need for endoscopic versus surgical treatment. Additional screening measures are directed toward the FAP-associated cancers listed in Table 13.2 . Unaffected individuals with wild-type APC who have an affected family member with a known mutation are screened in the same way as the general population.

Attenuated familial adenomatous polyposis

Clinical features.

Attenuated familial adenomatous polyposis (AFAP) is an FAP variant characterized by oligopolyposis (<100 colonic adenomas) and a CRC onset 10 to 20 years later than in patients with FAP, although the precise lifetime risk of CRC is not well defined. AFAP may display malignant and benign manifestations similar to those of FAP.

Genetics.

Like FAP, AFAP is an autosomal dominant cancer syndrome caused by germline mutations in the APC gene, with defects that tend to localize in the 3′ or 5′ regions of the gene.

You're Reading a Preview

Become a Clinical Tree membership for Full access and enjoy Unlimited articles

Become membership

If you are a member. Log in here