Entry - *191845 - UROMODULIN; UMOD - OMIM
 
* 191845

UROMODULIN; UMOD


Alternative titles; symbols

TAMM-HORSFALL GLYCOPROTEIN; THP; THGP


HGNC Approved Gene Symbol: UMOD

Cytogenetic location: 16p12.3     Genomic coordinates (GRCh38): 16:20,333,051-20,356,301 (from NCBI)


Gene-Phenotype Relationships
Location Phenotype Phenotype
MIM number
Inheritance Phenotype
mapping key
16p12.3 Tubulointerstitial kidney disease, autosomal dominant, 1 162000 AD 3

TEXT

Description

The Tamm-Horsfall glycoprotein (Tamm and Horsfall, 1952), also referred to as uromodulin (Muchmore and Decker, 1985), is a GPI-anchored glycoprotein and the most abundant protein in normal urine. Uromodulin, uropontin, and nephrocalcin are the 3 known urinary glycoproteins that affect the formation of calcium-containing kidney stones (167030). Uromodulin is synthesized by kidney and localizes in cells lining the ascending limb of Henle and distal convoluted tubule (McKenzie and McQueen, 1969; Schenk et al., 1971).


Cloning and Expression

Pennica et al. (1987) cloned uromodulin from human kidney total RNA. The deduced 639-amino acid protein contains an N-terminal signal peptide and regions of similarity with LDL receptor (LDLR; 606945) and the EGF precursor protein (131530). Uromodulin has 48 cysteines and 8 potential N-glycosylation sites. Northern blot analysis of about 150 different epithelial, connective, and hematopoietic tissues and tumor-derived cell lines detected a 2.6-kb transcript in adult kidney only.

Yang et al. (2004) described the domain structure of the UMOD protein. It has an N-terminal signal sequence, followed by 3 EGF-like repeats, a domain with 8 conserved cysteines (D8C), and a C-terminal zona pellucida (ZP) domain. The D8C domain is about 130 amino acids long, and the 8 cysteines are predicted to form 4 pairs of disulfide bridges. The secondary structure is composed mainly of beta strands.


Gene Structure

Pennica et al. (1987) reported that UMOD gene contains 11 exons.

Hart et al. (2002) subsequently determined that the UMOD gene contains 12 exons, with the novel exon being exon 2. Exons 1 and 2 are noncoding; the ATG translation initiation site is located in exon 3.


Mapping

Jeanpierre et al. (1993) mapped the UMOD gene to chromosome 16p13.11 by somatic cell hybrid analysis. By study of a different mapping panel, Pook et al. (1993) mapped the gene to chromosome 16p12.3-qter. Combining the results of the 2 studies suggested that the gene is located in the region 16p13.11-p12.3.

Fukuoka and Matsuda (1997) assigned the Umod gene to mouse chromosome 7F1-F2 and rat chromosome 1q36-q37 by fluorescence in situ hybridization. The result was discordant with the previous localization of the gene to mouse chromosome 17 by somatic cell hybrid analysis (Deng et al., 1995).


Gene Function

Muchmore and Decker (1985) purified 85-kD uromodulin from urine of pregnant women and found that it showed immunoregulatory potential. Uromodulin had a broad dose-response curve in inhibition of antigen-specific T-cell proliferation. It also inhibited monocyte reactivity in vitro and spontaneous monocyte cytotoxicity. It had no effect on B-cell function or cell viability.

Greimel et al. (2006) stated that approximately 30% of the mass of THP consists of sulfated N-linked glycans. They showed that both membrane-bound and soluble recombinant human GP3ST (GAL3ST2; 608237) incorporated radiolabeled SO4 into purified human urine THP in a concentration-dependent manner.

Using immunohistochemical studies, Dahan et al. (2003) found that uromodulin is distributed primarily at the apical membrane of the thick ascending loop and distal convoluted tubule in the human kidney. Uromodulin is a GPI-anchor-linked protein. It has putative roles in cell adhesion, signal transduction, inhibition of calcium oxalate crystal aggregation, defense against urinary tract infection, and modulation of urine-concentrating ability. It may also act as a potential nephritogenic antigen.

Zaucke et al. (2010) demonstrated that UMOD was expressed in the primary cilia of renal tubules. Immunofluorescent and ultrastructural studies confirmed ciliary expression of UMOD, with localization to the mitotic spindle poles and colocalization with ciliary proteins nephrocystin-1 (NPHP1; 607100) and kinesin family member 3A (KIF3A; 604683).

Using cryoelectron tomography, Weiss et al. (2020) showed that the human uromodulin filament consisted of a zigzag-shaped backbone with laterally protruding arms. N-glycosylation mapping and biophysical assays revealed that uromodulin acted as a multivalent ligand for the bacterial type-1 pilus adhesin, presenting specific epitopes on the regularly spaced arms. Imaging of uromodulin-uropathogen interactions in vitro and in patient urine showed that uromodulin filaments associated with uropathogens and mediated bacterial aggregation, likely preventing adhesion and allowing clearance by micturition.


Molecular Genetics

In affected members of 4 unrelated families with autosomal dominant tubulointerstitial kidney disease-1 (ADTKD1; 162000), Hart et al. (2002) identified 4 different heterozygous mutations in exon 4 of the UMOD gene (191845.0001-191845.0004). The mutations, which were found by a combination of linkage analysis and candidate gene sequencing, segregated with the disorder in the families. Three families had a clinical diagnosis of familial juvenile hyperuricemic nephropathy (FJHN, HNFJ), whereas the fourth family was diagnosed clinically with medullary cystic kidney disease (MCKD). In all families, the phenotype was characterized mainly by juvenile onset of hyperuricemia, polyuria, gout, and progressive renal insufficiency that was tubulointerstitial in origin. Renal biopsies showed tubular atrophy and interstitial fibrosis. Global glomerulosclerosis was also observed, although there was no evidence of glomerulonephritis. Necroscopy showed sheathing of the renal tubules by dense acellular hyaline fibrous tissue that likely represented abnormal deposition of the UMOD protein. Medullary cysts were present in 1 family. The authors noted that the disorder is associated with impaired urinary concentrating ability, which likely causes increased proximal tubular reabsorption of uric acid and hyperuricemia. Functional studies of the variants and studies of patient cells were not performed, but the authors postulated that the mutations caused tertiary structural changes in the uromodulin protein that could alter cytokine binding and ultimately lead to fibrosis and progressive renal failure. The report established that the clinical entities of FJHN and MCKD not only share clinical features, but are also either allelic or variable manifestations of the same disease. Noting that hyperuricemia and medullary cysts are variable features and that the conditions result from mutations in the same gene, the authors suggested the designation 'uromodulin-associated kidney disease.'

In 5 unrelated kindreds with ADTKD1, 2 from Austria and 3 from Spain, Turner et al. (2003) identified 5 heterozygous missense mutations in the UMOD gene (191845.0005-191845.0009) that altered evolutionary conserved residues. These mutations were not found in 110 alleles from 55 unrelated normal individuals. Functional studies of the variants were not performed, but the authors postulated a loss-of-function effect. The families had previously been reported by Stacey et al. (2003).

In affected members of 4 unrelated Italian families with variable manifestations of ADTKD1, Rampoldi et al. (2003) identified heterozygous missense mutations in the UMOD gene (see, e.g., C315R, 191845.0010 and C148W, 191845.0015). All mutations affected highly conserved cysteine residues and were predicted to affect protein structure. Immunohistochemistry of kidney biopsies revealed dense intracellular accumulation of uromodulin in tubular epithelia of the thick ascending limb of Henle loop. Electron microscopy showed accumulation of dense fibrillar material within the endoplasmic reticulum (ER), and patient urine samples consistently showed a severe reduction of excreted uromodulin. Experiments in transfected cells showed that all 4 mutations caused a delay in protein export to the plasma membrane due to a longer retention time in the ER. The protein maturation impairment and retention in the ER, which may trigger ubiquitination and ER stress, suggested a pathogenetic mechanism leading to these kidney diseases. Rampoldi et al. (2003) postulated that hyperuricemia is a secondary effect of volume contraction resulting from UMOD dysfunction in the thick ascending loop of Henle. Three families had a clinical diagnosis of MCKD/FJHN (including a family previously reported by Scolari et al., 1999), and 1 family had a clinical diagnosis of glomerulocystic kidney disease (GCKD), thus demonstrating that these clinical entities are allelic and are different manifestations of the same disorder.

In patients from 11 families with ADTKD1, Dahan et al. (2003) identified 11 different heterozygous UMOD mutations, including 10 novel ones (see, e.g., 191845.0012). All of the mutations occurred at highly conserved residues in exon 4, and 5 of the mutations affected a conserved cysteine residue. The families were ascertained from a larger group of 25 families with a similar phenotype; thus, UMOD mutations were found in 44% of families. Patient kidney samples showed abnormal uromodulin immunostaining within enlarged or cystic profiles within tubules in the thick ascending loop, and not at the apical membrane as observed in controls. Mutant UMOD was not found in proximal tubules. Patients also showed decreased urinary excretion of wildtype uromodulin. The findings indicated that mutant uromodulin accumulates within renal tubular cells in patients with UMOD mutations.

In affected members of 4 unrelated Spanish families with variable clinical manifestations of ADTKD1, Lens et al. (2005) identified heterozygous missense mutations in the UMOD gene (see, e.g., C300G, 191845.0009 and Q316P, 191845.0014). The mutations, which were found by a combination of linkage analysis and candidate gene sequencing, segregated with the disorder in available family members. Two families (F1 and F2) with the same mutation (Q316P) had clinical phenotypes consistent with MCKD (F1) and FJHN (F2). Family 4 had a phenotype consistent with FJHN, and F3, with the C300G mutation, had a phenotype consistent with GCKD. These genetic findings further supported the idea that MCKD, FJHN, and GCKD represent the same disease entity.

Vylet'al et al. (2006) sequenced the UMOD gene in 19 families with variable manifestations of ADTKD1 and identified heterozygous mutations in 6 families, 5 of which had been previously reported (kindred 6 from McBride et al., 1998; kindreds A and B from Stiburkova et al., 2000; Fairbanks et al., 2002; and family BE2 from Stiburkova et al., 2003) (see, e.g., 191845.0006 and 191845.0011). Functional studies of mutant proteins showed distinct glycosylation patterns, impaired intracellular trafficking, and decreased ability to be exposed on the plasma membrane, which correlated with observations in kidney tissue from patients. A reduction in urinary modulin excretion was found in 18 of the 19 families and was associated with case-specific differences in uromodulin immunohistochemical staining patterns in kidney tissues.

In 6 of 17 probands with ADTKD who were studied for UMOD mutations, Williams et al. (2009) identified 6 different heterozygous missense mutations in the UMOD gene (see, e.g., D196N; 191845.0016). In vitro functional studies of some of the mutations expressed in HeLa cells showed that the mutant uromodulins had significantly delayed maturation compared to wildtype, with abnormal protein retention in the ER and reduced or absent expression at the plasma membrane. There were different effects allowing the identification of 2 mutation groups: group A (including mutants C32W, D196N, and G488R) had 50% maturation compared to wildtype with some expression at the plasma membrane, whereas group B (including mutants C126R, 191845.0006; N128S, 191845.0007; and C223R) had 25% maturation compared to wildtype and absence of expression at the plasma membrane. There were no phenotypic differences between patients with group A and group B mutations. The findings suggested that abnormal folding of the mutant proteins resulted in protein retention in the ER, which may trigger apoptosis and underlie the mechanism for disease pathogenesis.

In affected members of 10 unrelated families with ADTKD1, Zaucke et al. (2010) identified 7 novel and 3 previously reported heterozygous missense mutations in the UMOD gene (see, e.g., 191845.0009 and 191845.0013). Most of the mutations affected conserved cysteine residues. The number of UMOD-positive primary cilia in renal biopsy samples from 2 of the patients was significantly decreased compared to control samples. The authors suggested that this defect may contribute to cyst formation. The families were ascertained from a cohort of 44 families with nephropathy from western Europe and the United States who underwent direct sequencing of the UMOD gene.

Using transfected MDCK cells, Ma et al. (2012) found that mutant mouse Thp with substitutions corresponding to human cys126 to arg (C126R; 191845.0006) or cys217 to gly (C217G; 191845.0012) resulted in protein misfolding, retention of mutant protein in the ER, and reduced surface Thp expression and secretion. In all measures, the C217G substitution within the D8C domain was more severe than the C126R substitution within EGF-like domain-3. Both substitutions also led to reduced expression of the chaperone protein Hsp70 (see HSPA1A; 140550), trapped wildtype Thp in the ER in a dominant-negative manner, and induced apoptosis. Treatments that favored correct protein folding tended to reduce the effect of the mutations. Greatest rescue was obtained with exposure to sodium 4-phenylbutyrate, a chemical chaperone that increases the expression of heat-shock proteins, and probenecid, a drug used clinically to treat hyperuricemia.


Animal Model

Type 1-fimbriated E. coli, which contains a mannose-sensitive lectin subunit at the fimbrial tip, binds THP, and this binding inhibits binding of E. coli to kidney epithelial cells. Mo et al. (2004) found that Thp -/- mice, which were obtained in the expected mendelian ratio, showed increased susceptibility to bladder infection by type 1-fimbriated E. coli.

Mo et al. (2004) found spontaneous formation of calcium crystals in kidneys of adult Thp -/- mice. Excessive intake of calcium and oxalate dramatically increased both the frequency and severity of renal calcium crystal formation in Thp -/- mice. High calcium/oxalate also induced Thp -/- renal epithelial cells to express osteopontin (OPN, or SPP1; 166490), an inhibitor of bone mineralization. Mo et al. (2004) hypothesized that OPN may be an inducible inhibitor of calcium crystallization, whereas THP is a constitutive inhibitor of calcium crystallization.

Bachmann et al. (2005) found that kidneys of Thp -/- mice were anatomically normal. However, Thp -/- kidneys showed reduced creatinine clearance, impaired water reabsorption following deprivation, and altered expression of renal transporters, channels, and regulatory molecules.

Bernascone et al. (2010) generated a transgenic mouse model with a Umod C147W mutation corresponding to the human UMOD C148W mutation (191845.0015). The mutant mice showed tubulointerstitial fibrosis with inflammatory cell infiltration, tubule dilation, and selective damage to the epithelial cells lining the thick ascending limb of the loop of Henle (TAL), leading to mild renal failure. Umod was retained in the ER of expressing cells, leading to ER hyperplasia. The authors suggested that impaired TAL function is a consequence of a gain-of-function effect of UMOD mutations.


ALLELIC VARIANTS ( 16 Selected Examples):

.0001 TUBULOINTERSTITIAL KIDNEY DISEASE, AUTOSOMAL DOMINANT, 1

UMOD, 27-BP DEL, NT529
  
RCV000518183...

In 36 individuals from a large multigenerational kindred (family 1) with autosomal dominant tubulointerstitial kidney disease-1 (ADTKD1; 162000), Hart et al. (2002) identified a heterozygous in-frame 27-bp deletion (c.529_555del) in exon 4 of the UMOD gene, resulting in the deletion of 9 amino acids (His177_Arg185del). The mutation, which was found by a combination of linkage analysis and candidate gene sequencing, segregated with the disorder in the family. The family contained more than 300 individuals spanning 7 generations. Four mutation carriers, who were all women, had normal serum uric acid despite low fractional excretions of uric acid; 2 had mild renal insufficiency. Functional studies of the variant and studies of patient cells were not performed. The original clinical diagnosis was familial juvenile hyperuricemic nephropathy (FJHN, HNFJ).


.0002 TUBULOINTERSTITIAL KIDNEY DISEASE, AUTOSOMAL DOMINANT, 1

UMOD, CYS148TYR
  
RCV002251318...

In 9 affected members of a 5-generation family (family 2) with autosomal dominant tubulointerstitial kidney disease-1 (ADTKD1; 162000), Hart et al. (2002) identified a heterozygous c.443G-A transition in exon 4 of the UMOD gene, resulting in a cys148-to-tyr (C148Y) substitution at a conserved residue. The mutation, which was found by a combination of linkage analysis and candidate gene sequencing, segregated with the disorder in the family. Functional studies of the variant and studies of patient cells were not performed. The original clinical diagnosis was familial juvenile hyperuricemic nephropathy (FJHN, HNFJ).


.0003 TUBULOINTERSTITIAL KIDNEY DISEASE, AUTOSOMAL DOMINANT, 1

UMOD, CYS217ARG
  
RCV002251319

In an affected individual from a family (family 4) with autosomal dominant tubulointerstitial kidney disease-1 (ADTKD1; 162000), Hart et al. (2002) identified a heterozygous c.649T-C transition in exon 4 of the UMOD gene, resulting in a cys217-to-arg (C217R) substitution at a conserved residue. Functional studies of the variant and studies of patient cells were not performed. The family had previously been reported by Massari et al. (1980) as having a clinical phenotype consistent with familial juvenile hyperuricemic nephropathy (FJHN, HNFJ).

Yang et al. (2004) determined that cys217 is 1 of 8 cysteines within UMOD predicted to form 4 disulfide bridges. The C217R mutation likely disrupts one of these disulfide bonds.


.0004 TUBULOINTERSTITIAL KIDNEY DISEASE, AUTOSOMAL DOMINANT, 1

UMOD, GLY103CYS
  
RCV002251320...

In 3 affected members of a multigenerational family (family 3) with autosomal dominant tubulointerstitial kidney disease-1 (ADTKD1; 162000), Hart et al. (2002) identified a heterozygous c.307G-T transversion in exon 4 of the UMOD gene, resulting in a gly103-to-cys (G103C) substitution at a conserved residue. Functional studies of the variant and studies of patient cells were not performed. The patients, who had hyperuricemia and renal insufficiency, were diagnosed clinically with medullary cystic kidney disease (MCKD).


.0005 TUBULOINTERSTITIAL KIDNEY DISEASE, AUTOSOMAL DOMINANT, 1

UMOD, CYS77TYR
  
RCV002251321

In affected members of a family (11/00) from Austria with autosomal dominant tubulointerstitial kidney disease-1 (ADTKD1; 162000), Turner et al. (2003) identified a heterozygous c.335G-A transition in the UMOD gene that was predicted to result in a cys77-to-tyr (C77Y) substitution. Functional studies of the variant and studies of patient cells were not performed. The family had previously been reported by Stacey et al. (2003).


.0006 TUBULOINTERSTITIAL KIDNEY DISEASE, AUTOSOMAL DOMINANT, 1

UMOD, CYS126ARG
  
RCV002251322

In affected members of a family (13/00) from Austria with autosomal dominant tubulointerstitial kidney disease-1 (ADTKD1; 162000), Turner et al. (2003) identified a heterozygous c.481T-C transition in the UMOD gene that was predicted to result in a cys126-to-arg (C126R) substitution. Functional studies of the variant and studies of patient cells were not performed. The family had previously been reported by Stacey et al. (2003).

In affected members of a family with ADTKD1, originally reported by Fairbanks et al. (2002), Vylet'al et al. (2006) identified heterozygosity for the C126R mutation in the UMOD gene. A female patient who had onset of disease at age 31, with gout, hypertension, elevated creatinine and hyperuricemia, underwent kidney transplantation but died at 36 years of age. A 43-year-old woman and 40-year-old man, who had onset of hyperuricemia at age 19 years and 16 years, respectively, had no abnormalities on renal ultrasound and had not developed any other symptoms.

In HeLa cells transfected with C126R mutant uromodulin, Williams et al. (2009) found that the mutant protein had delayed maturation (about 25% of wildtype), was retained in the ER, and not trafficked to the plasma membrane, likely due to protein misfolding. The authors suggested that abnormal protein retention in the ER may trigger apoptosis and underlie the mechanism for the pathogenesis of the disorder.

Using transfected MDCK cells, Ma et al. (2012) found that mutant mouse Thp with substitutions corresponding to human C126R or cys217 to gly (C217G; 191845.0012) resulted in protein misfolding, retention of mutant protein in the ER, and reduced surface Thp expression and secretion. In all measures, the C217G substitution within the D8C domain was more severe than the C126R substitution within EGF-like domain-3. Both substitutions also led to reduced expression of the chaperone protein Hsp70 (see HSPA1A; 140550), trapped wildtype Thp in the ER in a dominant-negative manner, and induced apoptosis. Treatments that favored correct protein folding tended to reduce the effect of the mutations. Greatest rescue was obtained with exposure to sodium 4-phenylbutyrate, a chemical chaperone that increases the expression of heat-shock proteins, and probenecid, a drug used clinically to treat hyperuricemia.


.0007 TUBULOINTERSTITIAL KIDNEY DISEASE, AUTOSOMAL DOMINANT, 1

UMOD, ASN128SER
  
RCV002251323

In affected members of a family from Spain (1/96) with autosomal dominant tubulointerstitial kidney disease-1 (ADTKD1; 162000), Turner et al. (2003) identified a heterozygous c.488A-G transition in the UMOD gene that was predicted to result in an asn128-to-ser (N128S) substitution. Functional studies of the variant and studies of patient cells were not performed. The family had previously been reported by Stacey et al. (2003).

In HeLa cells transfected with N128S mutant uromodulin, Williams et al. (2009) found that the mutant protein had delayed maturation (about 25% of wildtype), was retained in the ER, and not trafficked to the plasma membrane, likely due to protein misfolding. The authors suggested that abnormal protein retention in the ER may trigger apoptosis and underlie the mechanism for the pathogenesis of the disorder.


.0008 TUBULOINTERSTITIAL KIDNEY DISEASE, AUTOSOMAL DOMINANT, 1

UMOD, CYS255TYR
  
RCV002251324

In affected members of a family from Spain (13/96) with autosomal dominant tubulointerstitial kidney disease-1 (ADTKD1; 162000), Turner et al. (2003) identified a heterozygous c.869G-A transition in the UMOD gene, resulting in a cys255-to-tyr (C255Y) substitution at a conserved residue. Functional studies of the variant and studies of patient cells were not performed. The family had previously been reported by Stacey et al. (2003).

In 14 affected members of a large consanguineous Spanish family with ADTKD1, Rezende-Lima et al. (2004) identified a C255Y mutation in the UMOD gene. These authors stated that the substitution resulted from a c.764G-A transition in exon 4 and that the mutation disrupted a light chain-binding domain. Eleven family members were heterozygous for the mutation, whereas 3 were homozygous. The homozygous individuals had earlier disease onset than the heterozygous individuals, but the report demonstrated that homozygosity for UMOD mutations is not lethal. The clinical phenotype was consistent with medullary cystic kidney disease (MCKD), although the authors noted that cyst formation may be nonspecific secondary effect. This family was also reported as family F4 by Lens et al. (2005).


.0009 TUBULOINTERSTITIAL KIDNEY DISEASE, AUTOSOMAL DOMINANT, 1

UMOD, CYS300GLY
  
RCV002251325

In affected members of a family from Spain (20/96) with autosomal dominant tubulointerstitial kidney disease-1 (ADTKD1; 162000), Turner et al. (2003) identified a heterozygous c.1003T-G transversion in the UMOD gene that was predicted to result in a cys300-to-gly (C300G) substitution. Functional studies of the variant and studies of patient cells were not performed. The family had previously been reported by Stacey et al. (2003).

In affected members of a Spanish family (F3) with ADTKD1, Lens et al. (2005) identified heterozygosity for the C300G mutation in exon 5 of the UMOD gene. The phenotype included glomerulocystic kidney disease (GCKD), thus expanding the spectrum of manifestations associated with this mutation.


.0010 TUBULOINTERSTITIAL KIDNEY DISEASE, AUTOSOMAL DOMINANT, 1

UMOD, CYS315ARG
  
RCV002251326

In affected members of an Italian family (GCKD1) with autosomal dominant tubulointerstitial kidney disease-1 (ADTKD1; 162000), Rampoldi et al. (2003) identified a c.943T-C transition (c.943T-C, NM_003361) in exon 5 of the UMOD gene, predicted to result in a cys315-to-arg (C315R) substitution. The family had previously been reported by Scolari et al. (1999) with a clinical diagnosis of glomerulocystic kidney disease (GCKD).


.0011 TUBULOINTERSTITIAL KIDNEY DISEASE, AUTOSOMAL DOMINANT, 1

UMOD, VAL273PHE
  
RCV001566754...

In 3 Belgian brothers with autosomal dominant tubulointerstitial kidney disease-1 (ADTKD1; 162000), originally reported as family BE2 by Stiburkova et al. (2003), Vylet'al et al. (2006) identified a heterozygous mutation in the UMOD gene resulting in a val273-to-phe (V273F) substitution. Functional studies showed that the V273F-mutant protein was retained in the endoplasmic reticulum and did not reach the plasma membrane.


.0012 TUBULOINTERSTITIAL KIDNEY DISEASE, AUTOSOMAL DOMINANT, 1

UMOD, CYS217GLY
  
RCV002251328

In a 36-year-old Belgian woman with autosomal dominant tubulointerstitial kidney disease-1 (ADTKD1; 162000), Dahan et al. (2003) identified a heterozygous c.754T-G transversion in exon 4 of the UMOD gene, resulting in a cys217-to-gly (C217G) substitution. The patient was in end-stage renal failure, and she had a history of gout, with onset at age 21 years. Her father, grandmother, and 1 sister were also affected. Dahan et al. (2003) erroneously reported this mutation as 794T-G, instead of 754T-G, in their paper. In addition, their numbering of the UMOD sequence began at nucleotide 1, rather than at the ATG start codon at nucleotide 106. Following the convention of numbering from the start codon, this mutation should be referred to as c.649T-G (Gross, 2013).

Using transfected MCKD cells, Ma et al. (2012) found that mutant mouse Thp with substitutions corresponding to human cys126 to arg (C126R; 191845.0006) or C217G resulted in protein misfolding, retention of mutant protein in the ER, and reduced surface Thp expression and secretion. In all measures, the C217G substitution within the D8C domain was more severe than the C126R substitution within EGF-like domain-3. Both substitutions also led to reduced expression of the chaperone protein Hsp70 (see HSPA1A; 140550), trapped wildtype Thp in the ER in a dominant-negative manner, and induced apoptosis. Treatments that favored correct protein folding tended to reduce the effect of the mutations. Greatest rescue was obtained with exposure to sodium 4-phenylbutyrate, a chemical chaperone that increases the expression of heat-shock proteins, and probenecid, a drug used clinically to treat hyperuricemia.


.0013 TUBULOINTERSTITIAL KIDNEY DISEASE, AUTOSOMAL DOMINANT, 1

UMOD, CYS248SER
  
RCV002251329

In 3 members of a 2-generation family (F6) with autosomal dominant tubulointerstitial kidney disease-1 (ADTKD1; 162000), Zaucke et al. (2010) identified a heterozygous c.743G-C transversion in the UMOD gene, resulting in a cys248-to-ser substitution (C248S) at a conserved residue in the D8C domain. The father exhibited hypertension, gout, congestive heart failure, varices, and an aortic aneurysm. Renal biopsy showed tubular atrophy, tubulointerstitial fibrosis, and a thickened tubular basement membrane.


.0014 TUBULOINTERSTITIAL KIDNEY DISEASE, AUTOSOMAL DOMINANT, 1

UMOD, GLN316PRO
  
RCV002251372

In affected members of 2 apparently unrelated Spanish families (F1 and F2) with autosomal dominant tubulointerstitial kidney disease-1 (ADTKD1; 162000), Lens et al. (2005) identified a heterozygous c.947A-C transversion in exon 5 of the UMOD gene, resulting in a gln316-to-pro (Q316P) substitution at a conserved residue. The mutation segregated with the phenotype in the families and was not found in 100 control chromosomes. The clinical phenotype in family 1 was consistent with medullary cystic kidney disease (MCKD), whereas the clinical phenotype in family 2 was consistent with familial juvenile hyperuricemic nephropathy (FJHN). The results demonstrated that these variable clinical manifestations are part of the same disease entity, referred to as ADTKD1. Although functional studies of the variant were not performed, the authors suggested that the mutation could alter the tertiary structure of UMOD, leading to defects in vesicular transport in the thick ascending loop of Henle.


.0015 TUBULOINTERSTITIAL KIDNEY DISEASE, AUTOSOMAL DOMINANT, 1

UMOD, CYS148TRP
  
RCV002251407

In affected members of an Italian family (MCKD9) with autosomal dominant tubulointerstitial kidney disease-1 (ADTKD1; 162000), Rampoldi et al. (2003) identified a heterozygous c.444T-G transversion (c.444T-G, NM_003361) in exon 4 of the UMOD gene, resulting in a cys148-to-trp (C148W) substitution at a conserved residue.

Bernascone et al. (2010) generated a transgenic mouse model with a Umod C147W mutation corresponding to the human UMOD C148W mutation. The mutant mice showed tubulointerstitial fibrosis with inflammatory cell infiltration, tubule dilation, and selective damage to the epithelial cells lining the thick ascending limb of the loop of Henle (TAL), leading to mild renal failure. Umod was retained in the ER of expressing cells, leading to ER hyperplasia. The authors suggested that impaired TAL function is a consequence of a gain-of-function effect of UMOD mutations.


.0016 TUBULOINTERSTITIAL KIDNEY DISEASE, AUTOSOMAL DOMINANT, 1

UMOD, ASP196ASN
  
RCV002251408

In 3 affected members of a family (family 5) with autosomal dominant tubulointerstitial kidney disease-1 (ADTKD1; 162000), Williams et al. (2009) identified a heterozygous c.586G-A transition in exon 3 of the UMOD gene, resulting in an asp196-to-asn (D196N) substitution in the cysteine-rich domain. In vitro functional expression studies in HeLa cells transfected with the mutation showed that it caused a 50% delay in maturation compared to wildtype. The mutant protein was retained in the ER and showed decreased expression at the plasma membrane. The findings suggested that abnormal protein retention in the ER may trigger apoptosis and underlie the mechanism for disease pathogenesis.


REFERENCES

  1. Bachmann, S., Mutig, K., Bates, J., Welker, P., Geist, B., Gross, V., Luft, F. C., Alenina, N., Bader, M., Thiele, B. J., Prasadan, K., Raffi, H. S., Kumar, S. Renal effects of Tamm-Horsfall protein (uromodulin) deficiency in mice. Am. J. Physiol. Renal Physiol. 288: F559-F567, 2005. [PubMed: 15522986, related citations] [Full Text]

  2. Bernascone, I., Janas, S., Ikehata, M., Trudu, M., Corbelli, A., Schaeffer, C., Rastaldi, M. P., Devuyst, O., Rampoldi, L. A transgenic mouse model for uromodulin-associated kidney diseases shows specific tubulo-interstitial damage, urinary concentrating defect and renal failure. Hum. Molec. Genet. 19: 2998-3010, 2010. [PubMed: 20472742, related citations] [Full Text]

  3. Dahan, K., Devuyst, O., Smaers, M., Vertommen, D., Loute, G., Poux, J. M., Viron, B., Jacquot, C., Gagnadoux, M. F., Chauveau, D., Buchler, M., Cochat, P., Cosyns, J. P., Mougenot, B., Rider, M. H., Antignac, C., Verellen-Dumoulin, C., Pirson, Y. A cluster of mutations in the UMOD gene causes familial juvenile hyperuricemic nephropathy with abnormal expression of uromodulin. J. Am. Soc. Nephrol. 14: 2883-2893, 2003. [PubMed: 14569098, related citations] [Full Text]

  4. Deng, Z., Johnson, K., Engleward, B. P., Lane, S., Callen, D. F., Samson, L. D., Davisson, M. T., Siciliano, M. J. New regions of conserved synteny and linkage between human chromosome 16p12-p13 and mouse chromosomes 16 and 11. (Abstract) Cytogenet. Cell Genet. 68: 180 only, 1995.

  5. Fairbanks, L. D., Cameron, J. S., Venkat-Raman, G., Rigden, S. P. A., Rees, L., Van't Hoff, W., Mansell, M., Pattison, J., Goldsmith, D. J. A., Simmonds, H. A. Early treatment with allopurinol in familial juvenile hyperuricaemic nephropathy (FJHN) ameliorates the long-term progression of renal disease. Quart. J. Med. 95: 597-607, 2002. [PubMed: 12205338, related citations] [Full Text]

  6. Fukuoka, S.-I., Matsuda, Y. Assignment of the Tamm-Horsfall protein/uromodulin gene (Umod) to mouse chromosome bands 7F1-F2 and rat chromosome bands 1q36-q37 by in situ hybridization. Cytogenet. Cell Genet. 79: 241-242, 1997. [PubMed: 9605864, related citations] [Full Text]

  7. Greimel, P., Jabs, S., Storch, S., Cherif, S., Honke, K., Braulke, T., Thiem, J. In vitro sulfation of N-acetyllactosaminide by soluble recombinant human beta-Gal-3-prime-sulfotransferase. Carbohydr. Res. 341: 918-924, 2006. [PubMed: 16516177, related citations] [Full Text]

  8. Gross, M. B. Personal Communication. Baltimore, Md. 2/11/2013.

  9. Hart, T. C., Gorry, M. C., Hart, P. S., Woodard, A. S., Shihabi, Z., Sandhu, J., Shirts, B., Xu, L., Zhu, H., Barmada, M. M., Bleyer, A. J. Mutations of the UMOD gene are responsible for medullary cystic kidney disease 2 and familial juvenile hyperuricaemic nephropathy. J. Med. Genet. 39: 882-892, 2002. [PubMed: 12471200, related citations] [Full Text]

  10. Jeanpierre, C., Whitmore, S. A., Austruy, E., Cohen-Salmon, M., Callen, D. F., Junien, C. Chromosomal assignment of the uromodulin gene (UMOD) to 16p13.11. Cytogenet. Cell Genet. 62: 185-187, 1993. [PubMed: 8382593, related citations] [Full Text]

  11. Lens, X. M., Banet, J. F., Outeda, P., Barrio-Lucia, V. A novel pattern of mutation in uromodulin disorders: autosomal dominant medullary cystic kidney disease type 2, familial juvenile hyperuricemic nephropathy, and autosomal dominant glomerulocystic kidney disease. Am. J. Kidney Dis. 46: 52-57, 2005. [PubMed: 15983957, related citations] [Full Text]

  12. Ma, L., Liu, Y., El-Achkar, T. M., Wu, X.-R. Molecular and cellular effects of Tamm-Horsfall protein mutations and their rescue by chemical chaperones. J. Biol. Chem. 287: 1290-1305, 2012. [PubMed: 22117067, images, related citations] [Full Text]

  13. Massari, P. U., Hsu, C. H., Barnes, R. V., Fox, I. H., Gikas, P. W., Weller, J. M. Familial hyperuricemia and renal disease. Arch. Intern. Med. 140: 680-684, 1980. [PubMed: 7396593, related citations]

  14. McBride, M. B., Ridgen, S., Haycock, G. B., Dalton, N., Van't Hoff, W., Rees, L., Venkat-Raman, G. V., Moro, F., Ogg, C. S., Cameron, J. S., Simmonds, H. A. Presymptomatic detection of familial juvenile hyperuricaemic nephropathy in children. Pediat. Nephrol. 12: 357-364, 1998. [PubMed: 9686952, related citations] [Full Text]

  15. McKenzie, J. K., McQueen, E. G. Immunofluorescent localization of Tamm-Horsfall mucoprotein in human kidney. J. Clin. Path. 22: 334-339, 1969. [PubMed: 4891482, related citations] [Full Text]

  16. Mo, L., Huang, H.-Y., Zhu, X.-H., Shapiro, E., Hasty, D. L., Wu, X.-R. Tamm-Horsfall protein is a critical renal defense factor protecting against calcium oxalate crystal formation. Kidney Int. 66: 1159-1166, 2004. [PubMed: 15327412, related citations] [Full Text]

  17. Mo, L., Zhu, X.-H., Huang, H.-Y., Shapiro, E., Hasty, D. L., Wu, X.-R. Ablation of the Tamm-Horsfall protein gene increases susceptibility of mice to bladder colonization by type 1-fimbriated Escherichia coli. Am. J. Physiol. Renal Physiol. 286: F795-F802, 2004. [PubMed: 14665435, related citations] [Full Text]

  18. Muchmore, A. V., Decker, J. M. Uromodulin: a unique 85-kilodalton immunosuppressive glycoprotein isolated from urine of pregnant women. Science 229: 479-481, 1985. [PubMed: 2409603, related citations] [Full Text]

  19. Pennica, D., Kohr, W. J., Kuang, W.-J., Glaister, D., Aggarwal, B. B., Chen, E. Y., Goeddel, D. V. Identification of human uromodulin as the Tamm-Horsfall urinary glycoprotein. Science 236: 83-88, 1987. [PubMed: 3453112, related citations] [Full Text]

  20. Pook, M. A., Jeremiah, S., Scheinman, S. J., Povey, S., Thakker, R. V. Localization of the Tamm-Horsfall glycoprotein (uromodulin) gene to chromosome 16p12.3-16p13.11. Ann. Hum. Genet. 57: 285-290, 1993. [PubMed: 8179291, related citations] [Full Text]

  21. Rampoldi, L., Caridi, G., Santon, D., Boaretto, F., Bernascone, I., Lamorte, G., Tardanico, R., Dagnino, M., Colussi, G., Scolari, F., Ghiggeri, G. M., Amoroso, A., Casari, G. Allelism of MCKD, FJHN and GCKD caused by impairment of uromodulin export dynamics. Hum. Molec. Genet. 12: 3369-3384, 2003. [PubMed: 14570709, related citations] [Full Text]

  22. Rezende-Lima, W., Parreira, K. S., Garcia-Gonzalez, M., Riveira, E., Banet, J. F., Lens, X. M. Homozygosity for uromodulin disorders: FJHN and MCKD-type 2. Kidney Int. 66: 558-563, 2004. [PubMed: 15253706, related citations] [Full Text]

  23. Schenk, E. A., Schwartz, R. H., Lewis, R. A. Tamm-Horsfall mucoprotein. I. Localization in the kidney. Lab Invest. 25: 92-95, 1971. [PubMed: 4933557, related citations]

  24. Scolari, F., Puzzer, D., Amoroso, A., Caridi, G., Ghiggeri, G. M., Maiorca, R., Aridon, P., De Fusco, M., Ballabio, A., Casari, G. Identification of a new locus for medullary cystic disease, on chromosome 16p12. Am. J. Hum. Genet. 64: 1655-1660, 1999. [PubMed: 10330352, related citations] [Full Text]

  25. Stacey, J. M., Turner, J. J. O., Harding, B., Nesbit, M. A., Kotanko, P., Lhotta, K., Puig, J. G., Torres, R. J., Thakker, R. V. Genetic mapping studies of familial juvenile hyperuricemic nephropathy on chromosome 16p11-p13. J. Clin. Endocr. Metab. 88: 464-470, 2003. [PubMed: 12519891, related citations] [Full Text]

  26. Stiburkova, B., Majewski, J., Hodanova, K., Ondrova, L., Jerabkova, M., Zikanova, M., Vylet'al, P., Sebesta, I., Marinaki, A., Simmonds, A., Matthijs, G., Fryns, J.-P., Torres, R., Puig, J. G., Ott, J., Kmoch, S. Familial juvenile hyperuricaemic nephropathy (FJHN): linkage analysis in 15 families, physical and transcriptional characterisation of the FJHN critical region on chromosome 16p11.2 and the analysis of seven candidate genes. Europ. J. Hum. Genet. 11: 145-154, 2003. [PubMed: 12634862, related citations] [Full Text]

  27. Stiburkova, B., Majewski, J., Sebesta, I., Zhang, W., Ott, J., Kmoch, S. Familial juvenile hyperuricemic nephropathy: localization of the gene on chromosome 16p11.2--and evidence for genetic heterogeneity. Am. J. Hum. Genet. 66: 1989-1994, 2000. [PubMed: 10780922, related citations] [Full Text]

  28. Tamm, I., Horsfall, F. L., Jr. A mucoprotein derived from human urine which reacts with influenza, mumps, and Newcastle disease viruses. J. Exp. Med. 95: 71-97, 1952. [PubMed: 14907962, related citations] [Full Text]

  29. Turner, J. J. O., Stacey, J. M., Harding, B., Kotanko, P., Lhotta, K., Puig, J. G., Roberts, I., Torres, R. J., Thakker, R. V. UROMODULIN mutations cause familial juvenile hyperuricemic nephropathy. J. Clin. Endocr. Metab. 88: 1398-1401, 2003. [PubMed: 12629136, related citations] [Full Text]

  30. Vylet'al, P., Kublova, M., Kalbacova, M., Hodanova, K., Baresova, V., Stiburkova, B., Sikora, J., Hulkova, H., Zivny, J., Majewski, J., Simmonds, A., Fryns, J.-P., Venkat-Raman, G., Elleder, M., Kmoch, S. Alterations of uromodulin biology: a common denominator of the genetically heterogeneous FJHN/MCKD syndrome. Kidney Int. 70: 1155-1169, 2006. [PubMed: 16883323, related citations] [Full Text]

  31. Weiss, G. L., Stanisich, J. J., Sauer, M. M., Lin, C.-W., Eras, J., Zyla, D. S., Truck, J., Devuyst, O., Aebi, M., Pilhofer, M., Glockshuber, R. Architecture and function of human uromodulin filaments in urinary tract infections. Science 369: 1005-1010, 2020. [PubMed: 32616672, related citations] [Full Text]

  32. Williams, S. E., Reed, A. A. C., Galvanovskis, J., Antignac, C., Goodship, T., Karet, F. E., Kotanko, P., Lhotta, K., Moriniere, V., Williams, P., Wong, W., Rorsman, P., Thakker, R. V. Uromodulin mutations causing familial juvenile hyperuricaemic nephropathy lead to protein maturation defects and retention in the endoplasmic reticulum. Hum. Molec. Genet. 18: 2963-2974, 2009. [PubMed: 19465746, images, related citations] [Full Text]

  33. Yang, H., Wu, C., Zhao, S., Guo, J. Identification and characterization of D8C, a novel domain present in liver-specific LZP, uromodulin and glycoprotein 2, mutated in familial juvenile hyperuricaemic nephropathy. FEBS Lett. 578: 236-238, 2004. [PubMed: 15589826, related citations] [Full Text]

  34. Zaucke, F., Boehnlein, J. M., Steffens, S., Polishchuk, R. S., Rampoldi, L., Fischer, A., Pasch, A., Boehm, C. W. A., Baasner, A., Attanasio, M., Hoppe, B., Hopfer, H., Beck, B. B., Sayer, J. A., Hildebrandt, F., Wolf, M. T. F. Uromodulin is expressed in renal primary cilia and UMOD mutations result in decreased ciliary uromodulin expression. Hum. Molec. Genet. 19: 1985-1997, 2010. [PubMed: 20172860, images, related citations] [Full Text]


Cassandra L. Kniffin - updated : 01/26/2021
Ada Hamosh - updated : 12/04/2020
Carol A. Bocchini - updated : 10/18/2017
George E. Tiller - updated : 9/9/2013
Cassandra L. Kniffin - updated : 2/13/2013
Matthew B. Gross - updated : 2/11/2013
Patricia A. Hartz - updated : 2/8/2013
George E. Tiller - updated : 6/25/2010
Marla J. F. O'Neill - updated : 10/15/2009
George E. Tiller - updated : 1/10/2006
Patricia A. Hartz - updated : 4/19/2005
John A. Phillips, III - updated : 3/29/2004
Victor A. McKusick - updated : 6/30/2003
Victor A. McKusick - updated : 5/28/1998
Creation Date:
Victor A. McKusick : 6/6/1994
alopez : 02/09/2021
ckniffin : 01/26/2021
mgross : 12/04/2020
carol : 10/19/2017
carol : 10/18/2017
alopez : 09/09/2013
alopez : 9/9/2013
tpirozzi : 9/5/2013
tpirozzi : 9/5/2013
carol : 3/4/2013
ckniffin : 2/13/2013
mgross : 2/11/2013
mgross : 2/11/2013
terry : 2/8/2013
mgross : 8/3/2011
carol : 12/10/2010
wwang : 7/15/2010
terry : 6/25/2010
alopez : 1/6/2010
terry : 10/21/2009
wwang : 10/16/2009
terry : 10/15/2009
carol : 6/18/2008
terry : 12/17/2007
wwang : 2/10/2006
terry : 1/10/2006
mgross : 4/25/2005
terry : 4/19/2005
alopez : 3/29/2004
tkritzer : 7/15/2003
tkritzer : 7/7/2003
terry : 6/30/2003
terry : 6/1/1998
terry : 5/28/1998
jason : 6/6/1994

* 191845

UROMODULIN; UMOD


Alternative titles; symbols

TAMM-HORSFALL GLYCOPROTEIN; THP; THGP


HGNC Approved Gene Symbol: UMOD

Cytogenetic location: 16p12.3     Genomic coordinates (GRCh38): 16:20,333,051-20,356,301 (from NCBI)


Gene-Phenotype Relationships

Location Phenotype Phenotype
MIM number
Inheritance Phenotype
mapping key
16p12.3 Tubulointerstitial kidney disease, autosomal dominant, 1 162000 Autosomal dominant 3

TEXT

Description

The Tamm-Horsfall glycoprotein (Tamm and Horsfall, 1952), also referred to as uromodulin (Muchmore and Decker, 1985), is a GPI-anchored glycoprotein and the most abundant protein in normal urine. Uromodulin, uropontin, and nephrocalcin are the 3 known urinary glycoproteins that affect the formation of calcium-containing kidney stones (167030). Uromodulin is synthesized by kidney and localizes in cells lining the ascending limb of Henle and distal convoluted tubule (McKenzie and McQueen, 1969; Schenk et al., 1971).


Cloning and Expression

Pennica et al. (1987) cloned uromodulin from human kidney total RNA. The deduced 639-amino acid protein contains an N-terminal signal peptide and regions of similarity with LDL receptor (LDLR; 606945) and the EGF precursor protein (131530). Uromodulin has 48 cysteines and 8 potential N-glycosylation sites. Northern blot analysis of about 150 different epithelial, connective, and hematopoietic tissues and tumor-derived cell lines detected a 2.6-kb transcript in adult kidney only.

Yang et al. (2004) described the domain structure of the UMOD protein. It has an N-terminal signal sequence, followed by 3 EGF-like repeats, a domain with 8 conserved cysteines (D8C), and a C-terminal zona pellucida (ZP) domain. The D8C domain is about 130 amino acids long, and the 8 cysteines are predicted to form 4 pairs of disulfide bridges. The secondary structure is composed mainly of beta strands.


Gene Structure

Pennica et al. (1987) reported that UMOD gene contains 11 exons.

Hart et al. (2002) subsequently determined that the UMOD gene contains 12 exons, with the novel exon being exon 2. Exons 1 and 2 are noncoding; the ATG translation initiation site is located in exon 3.


Mapping

Jeanpierre et al. (1993) mapped the UMOD gene to chromosome 16p13.11 by somatic cell hybrid analysis. By study of a different mapping panel, Pook et al. (1993) mapped the gene to chromosome 16p12.3-qter. Combining the results of the 2 studies suggested that the gene is located in the region 16p13.11-p12.3.

Fukuoka and Matsuda (1997) assigned the Umod gene to mouse chromosome 7F1-F2 and rat chromosome 1q36-q37 by fluorescence in situ hybridization. The result was discordant with the previous localization of the gene to mouse chromosome 17 by somatic cell hybrid analysis (Deng et al., 1995).


Gene Function

Muchmore and Decker (1985) purified 85-kD uromodulin from urine of pregnant women and found that it showed immunoregulatory potential. Uromodulin had a broad dose-response curve in inhibition of antigen-specific T-cell proliferation. It also inhibited monocyte reactivity in vitro and spontaneous monocyte cytotoxicity. It had no effect on B-cell function or cell viability.

Greimel et al. (2006) stated that approximately 30% of the mass of THP consists of sulfated N-linked glycans. They showed that both membrane-bound and soluble recombinant human GP3ST (GAL3ST2; 608237) incorporated radiolabeled SO4 into purified human urine THP in a concentration-dependent manner.

Using immunohistochemical studies, Dahan et al. (2003) found that uromodulin is distributed primarily at the apical membrane of the thick ascending loop and distal convoluted tubule in the human kidney. Uromodulin is a GPI-anchor-linked protein. It has putative roles in cell adhesion, signal transduction, inhibition of calcium oxalate crystal aggregation, defense against urinary tract infection, and modulation of urine-concentrating ability. It may also act as a potential nephritogenic antigen.

Zaucke et al. (2010) demonstrated that UMOD was expressed in the primary cilia of renal tubules. Immunofluorescent and ultrastructural studies confirmed ciliary expression of UMOD, with localization to the mitotic spindle poles and colocalization with ciliary proteins nephrocystin-1 (NPHP1; 607100) and kinesin family member 3A (KIF3A; 604683).

Using cryoelectron tomography, Weiss et al. (2020) showed that the human uromodulin filament consisted of a zigzag-shaped backbone with laterally protruding arms. N-glycosylation mapping and biophysical assays revealed that uromodulin acted as a multivalent ligand for the bacterial type-1 pilus adhesin, presenting specific epitopes on the regularly spaced arms. Imaging of uromodulin-uropathogen interactions in vitro and in patient urine showed that uromodulin filaments associated with uropathogens and mediated bacterial aggregation, likely preventing adhesion and allowing clearance by micturition.


Molecular Genetics

In affected members of 4 unrelated families with autosomal dominant tubulointerstitial kidney disease-1 (ADTKD1; 162000), Hart et al. (2002) identified 4 different heterozygous mutations in exon 4 of the UMOD gene (191845.0001-191845.0004). The mutations, which were found by a combination of linkage analysis and candidate gene sequencing, segregated with the disorder in the families. Three families had a clinical diagnosis of familial juvenile hyperuricemic nephropathy (FJHN, HNFJ), whereas the fourth family was diagnosed clinically with medullary cystic kidney disease (MCKD). In all families, the phenotype was characterized mainly by juvenile onset of hyperuricemia, polyuria, gout, and progressive renal insufficiency that was tubulointerstitial in origin. Renal biopsies showed tubular atrophy and interstitial fibrosis. Global glomerulosclerosis was also observed, although there was no evidence of glomerulonephritis. Necroscopy showed sheathing of the renal tubules by dense acellular hyaline fibrous tissue that likely represented abnormal deposition of the UMOD protein. Medullary cysts were present in 1 family. The authors noted that the disorder is associated with impaired urinary concentrating ability, which likely causes increased proximal tubular reabsorption of uric acid and hyperuricemia. Functional studies of the variants and studies of patient cells were not performed, but the authors postulated that the mutations caused tertiary structural changes in the uromodulin protein that could alter cytokine binding and ultimately lead to fibrosis and progressive renal failure. The report established that the clinical entities of FJHN and MCKD not only share clinical features, but are also either allelic or variable manifestations of the same disease. Noting that hyperuricemia and medullary cysts are variable features and that the conditions result from mutations in the same gene, the authors suggested the designation 'uromodulin-associated kidney disease.'

In 5 unrelated kindreds with ADTKD1, 2 from Austria and 3 from Spain, Turner et al. (2003) identified 5 heterozygous missense mutations in the UMOD gene (191845.0005-191845.0009) that altered evolutionary conserved residues. These mutations were not found in 110 alleles from 55 unrelated normal individuals. Functional studies of the variants were not performed, but the authors postulated a loss-of-function effect. The families had previously been reported by Stacey et al. (2003).

In affected members of 4 unrelated Italian families with variable manifestations of ADTKD1, Rampoldi et al. (2003) identified heterozygous missense mutations in the UMOD gene (see, e.g., C315R, 191845.0010 and C148W, 191845.0015). All mutations affected highly conserved cysteine residues and were predicted to affect protein structure. Immunohistochemistry of kidney biopsies revealed dense intracellular accumulation of uromodulin in tubular epithelia of the thick ascending limb of Henle loop. Electron microscopy showed accumulation of dense fibrillar material within the endoplasmic reticulum (ER), and patient urine samples consistently showed a severe reduction of excreted uromodulin. Experiments in transfected cells showed that all 4 mutations caused a delay in protein export to the plasma membrane due to a longer retention time in the ER. The protein maturation impairment and retention in the ER, which may trigger ubiquitination and ER stress, suggested a pathogenetic mechanism leading to these kidney diseases. Rampoldi et al. (2003) postulated that hyperuricemia is a secondary effect of volume contraction resulting from UMOD dysfunction in the thick ascending loop of Henle. Three families had a clinical diagnosis of MCKD/FJHN (including a family previously reported by Scolari et al., 1999), and 1 family had a clinical diagnosis of glomerulocystic kidney disease (GCKD), thus demonstrating that these clinical entities are allelic and are different manifestations of the same disorder.

In patients from 11 families with ADTKD1, Dahan et al. (2003) identified 11 different heterozygous UMOD mutations, including 10 novel ones (see, e.g., 191845.0012). All of the mutations occurred at highly conserved residues in exon 4, and 5 of the mutations affected a conserved cysteine residue. The families were ascertained from a larger group of 25 families with a similar phenotype; thus, UMOD mutations were found in 44% of families. Patient kidney samples showed abnormal uromodulin immunostaining within enlarged or cystic profiles within tubules in the thick ascending loop, and not at the apical membrane as observed in controls. Mutant UMOD was not found in proximal tubules. Patients also showed decreased urinary excretion of wildtype uromodulin. The findings indicated that mutant uromodulin accumulates within renal tubular cells in patients with UMOD mutations.

In affected members of 4 unrelated Spanish families with variable clinical manifestations of ADTKD1, Lens et al. (2005) identified heterozygous missense mutations in the UMOD gene (see, e.g., C300G, 191845.0009 and Q316P, 191845.0014). The mutations, which were found by a combination of linkage analysis and candidate gene sequencing, segregated with the disorder in available family members. Two families (F1 and F2) with the same mutation (Q316P) had clinical phenotypes consistent with MCKD (F1) and FJHN (F2). Family 4 had a phenotype consistent with FJHN, and F3, with the C300G mutation, had a phenotype consistent with GCKD. These genetic findings further supported the idea that MCKD, FJHN, and GCKD represent the same disease entity.

Vylet'al et al. (2006) sequenced the UMOD gene in 19 families with variable manifestations of ADTKD1 and identified heterozygous mutations in 6 families, 5 of which had been previously reported (kindred 6 from McBride et al., 1998; kindreds A and B from Stiburkova et al., 2000; Fairbanks et al., 2002; and family BE2 from Stiburkova et al., 2003) (see, e.g., 191845.0006 and 191845.0011). Functional studies of mutant proteins showed distinct glycosylation patterns, impaired intracellular trafficking, and decreased ability to be exposed on the plasma membrane, which correlated with observations in kidney tissue from patients. A reduction in urinary modulin excretion was found in 18 of the 19 families and was associated with case-specific differences in uromodulin immunohistochemical staining patterns in kidney tissues.

In 6 of 17 probands with ADTKD who were studied for UMOD mutations, Williams et al. (2009) identified 6 different heterozygous missense mutations in the UMOD gene (see, e.g., D196N; 191845.0016). In vitro functional studies of some of the mutations expressed in HeLa cells showed that the mutant uromodulins had significantly delayed maturation compared to wildtype, with abnormal protein retention in the ER and reduced or absent expression at the plasma membrane. There were different effects allowing the identification of 2 mutation groups: group A (including mutants C32W, D196N, and G488R) had 50% maturation compared to wildtype with some expression at the plasma membrane, whereas group B (including mutants C126R, 191845.0006; N128S, 191845.0007; and C223R) had 25% maturation compared to wildtype and absence of expression at the plasma membrane. There were no phenotypic differences between patients with group A and group B mutations. The findings suggested that abnormal folding of the mutant proteins resulted in protein retention in the ER, which may trigger apoptosis and underlie the mechanism for disease pathogenesis.

In affected members of 10 unrelated families with ADTKD1, Zaucke et al. (2010) identified 7 novel and 3 previously reported heterozygous missense mutations in the UMOD gene (see, e.g., 191845.0009 and 191845.0013). Most of the mutations affected conserved cysteine residues. The number of UMOD-positive primary cilia in renal biopsy samples from 2 of the patients was significantly decreased compared to control samples. The authors suggested that this defect may contribute to cyst formation. The families were ascertained from a cohort of 44 families with nephropathy from western Europe and the United States who underwent direct sequencing of the UMOD gene.

Using transfected MDCK cells, Ma et al. (2012) found that mutant mouse Thp with substitutions corresponding to human cys126 to arg (C126R; 191845.0006) or cys217 to gly (C217G; 191845.0012) resulted in protein misfolding, retention of mutant protein in the ER, and reduced surface Thp expression and secretion. In all measures, the C217G substitution within the D8C domain was more severe than the C126R substitution within EGF-like domain-3. Both substitutions also led to reduced expression of the chaperone protein Hsp70 (see HSPA1A; 140550), trapped wildtype Thp in the ER in a dominant-negative manner, and induced apoptosis. Treatments that favored correct protein folding tended to reduce the effect of the mutations. Greatest rescue was obtained with exposure to sodium 4-phenylbutyrate, a chemical chaperone that increases the expression of heat-shock proteins, and probenecid, a drug used clinically to treat hyperuricemia.


Animal Model

Type 1-fimbriated E. coli, which contains a mannose-sensitive lectin subunit at the fimbrial tip, binds THP, and this binding inhibits binding of E. coli to kidney epithelial cells. Mo et al. (2004) found that Thp -/- mice, which were obtained in the expected mendelian ratio, showed increased susceptibility to bladder infection by type 1-fimbriated E. coli.

Mo et al. (2004) found spontaneous formation of calcium crystals in kidneys of adult Thp -/- mice. Excessive intake of calcium and oxalate dramatically increased both the frequency and severity of renal calcium crystal formation in Thp -/- mice. High calcium/oxalate also induced Thp -/- renal epithelial cells to express osteopontin (OPN, or SPP1; 166490), an inhibitor of bone mineralization. Mo et al. (2004) hypothesized that OPN may be an inducible inhibitor of calcium crystallization, whereas THP is a constitutive inhibitor of calcium crystallization.

Bachmann et al. (2005) found that kidneys of Thp -/- mice were anatomically normal. However, Thp -/- kidneys showed reduced creatinine clearance, impaired water reabsorption following deprivation, and altered expression of renal transporters, channels, and regulatory molecules.

Bernascone et al. (2010) generated a transgenic mouse model with a Umod C147W mutation corresponding to the human UMOD C148W mutation (191845.0015). The mutant mice showed tubulointerstitial fibrosis with inflammatory cell infiltration, tubule dilation, and selective damage to the epithelial cells lining the thick ascending limb of the loop of Henle (TAL), leading to mild renal failure. Umod was retained in the ER of expressing cells, leading to ER hyperplasia. The authors suggested that impaired TAL function is a consequence of a gain-of-function effect of UMOD mutations.


ALLELIC VARIANTS 16 Selected Examples):

.0001   TUBULOINTERSTITIAL KIDNEY DISEASE, AUTOSOMAL DOMINANT, 1

UMOD, 27-BP DEL, NT529
SNP: rs1555487528, ClinVar: RCV000518183, RCV002251317

In 36 individuals from a large multigenerational kindred (family 1) with autosomal dominant tubulointerstitial kidney disease-1 (ADTKD1; 162000), Hart et al. (2002) identified a heterozygous in-frame 27-bp deletion (c.529_555del) in exon 4 of the UMOD gene, resulting in the deletion of 9 amino acids (His177_Arg185del). The mutation, which was found by a combination of linkage analysis and candidate gene sequencing, segregated with the disorder in the family. The family contained more than 300 individuals spanning 7 generations. Four mutation carriers, who were all women, had normal serum uric acid despite low fractional excretions of uric acid; 2 had mild renal insufficiency. Functional studies of the variant and studies of patient cells were not performed. The original clinical diagnosis was familial juvenile hyperuricemic nephropathy (FJHN, HNFJ).


.0002   TUBULOINTERSTITIAL KIDNEY DISEASE, AUTOSOMAL DOMINANT, 1

UMOD, CYS148TYR
SNP: rs28934582, ClinVar: RCV002251318, RCV002513003

In 9 affected members of a 5-generation family (family 2) with autosomal dominant tubulointerstitial kidney disease-1 (ADTKD1; 162000), Hart et al. (2002) identified a heterozygous c.443G-A transition in exon 4 of the UMOD gene, resulting in a cys148-to-tyr (C148Y) substitution at a conserved residue. The mutation, which was found by a combination of linkage analysis and candidate gene sequencing, segregated with the disorder in the family. Functional studies of the variant and studies of patient cells were not performed. The original clinical diagnosis was familial juvenile hyperuricemic nephropathy (FJHN, HNFJ).


.0003   TUBULOINTERSTITIAL KIDNEY DISEASE, AUTOSOMAL DOMINANT, 1

UMOD, CYS217ARG
SNP: rs28934583, ClinVar: RCV002251319

In an affected individual from a family (family 4) with autosomal dominant tubulointerstitial kidney disease-1 (ADTKD1; 162000), Hart et al. (2002) identified a heterozygous c.649T-C transition in exon 4 of the UMOD gene, resulting in a cys217-to-arg (C217R) substitution at a conserved residue. Functional studies of the variant and studies of patient cells were not performed. The family had previously been reported by Massari et al. (1980) as having a clinical phenotype consistent with familial juvenile hyperuricemic nephropathy (FJHN, HNFJ).

Yang et al. (2004) determined that cys217 is 1 of 8 cysteines within UMOD predicted to form 4 disulfide bridges. The C217R mutation likely disrupts one of these disulfide bonds.


.0004   TUBULOINTERSTITIAL KIDNEY DISEASE, AUTOSOMAL DOMINANT, 1

UMOD, GLY103CYS
SNP: rs28934584, gnomAD: rs28934584, ClinVar: RCV002251320, RCV003398493

In 3 affected members of a multigenerational family (family 3) with autosomal dominant tubulointerstitial kidney disease-1 (ADTKD1; 162000), Hart et al. (2002) identified a heterozygous c.307G-T transversion in exon 4 of the UMOD gene, resulting in a gly103-to-cys (G103C) substitution at a conserved residue. Functional studies of the variant and studies of patient cells were not performed. The patients, who had hyperuricemia and renal insufficiency, were diagnosed clinically with medullary cystic kidney disease (MCKD).


.0005   TUBULOINTERSTITIAL KIDNEY DISEASE, AUTOSOMAL DOMINANT, 1

UMOD, CYS77TYR
SNP: rs121917768, ClinVar: RCV002251321

In affected members of a family (11/00) from Austria with autosomal dominant tubulointerstitial kidney disease-1 (ADTKD1; 162000), Turner et al. (2003) identified a heterozygous c.335G-A transition in the UMOD gene that was predicted to result in a cys77-to-tyr (C77Y) substitution. Functional studies of the variant and studies of patient cells were not performed. The family had previously been reported by Stacey et al. (2003).


.0006   TUBULOINTERSTITIAL KIDNEY DISEASE, AUTOSOMAL DOMINANT, 1

UMOD, CYS126ARG
SNP: rs121917769, ClinVar: RCV002251322

In affected members of a family (13/00) from Austria with autosomal dominant tubulointerstitial kidney disease-1 (ADTKD1; 162000), Turner et al. (2003) identified a heterozygous c.481T-C transition in the UMOD gene that was predicted to result in a cys126-to-arg (C126R) substitution. Functional studies of the variant and studies of patient cells were not performed. The family had previously been reported by Stacey et al. (2003).

In affected members of a family with ADTKD1, originally reported by Fairbanks et al. (2002), Vylet'al et al. (2006) identified heterozygosity for the C126R mutation in the UMOD gene. A female patient who had onset of disease at age 31, with gout, hypertension, elevated creatinine and hyperuricemia, underwent kidney transplantation but died at 36 years of age. A 43-year-old woman and 40-year-old man, who had onset of hyperuricemia at age 19 years and 16 years, respectively, had no abnormalities on renal ultrasound and had not developed any other symptoms.

In HeLa cells transfected with C126R mutant uromodulin, Williams et al. (2009) found that the mutant protein had delayed maturation (about 25% of wildtype), was retained in the ER, and not trafficked to the plasma membrane, likely due to protein misfolding. The authors suggested that abnormal protein retention in the ER may trigger apoptosis and underlie the mechanism for the pathogenesis of the disorder.

Using transfected MDCK cells, Ma et al. (2012) found that mutant mouse Thp with substitutions corresponding to human C126R or cys217 to gly (C217G; 191845.0012) resulted in protein misfolding, retention of mutant protein in the ER, and reduced surface Thp expression and secretion. In all measures, the C217G substitution within the D8C domain was more severe than the C126R substitution within EGF-like domain-3. Both substitutions also led to reduced expression of the chaperone protein Hsp70 (see HSPA1A; 140550), trapped wildtype Thp in the ER in a dominant-negative manner, and induced apoptosis. Treatments that favored correct protein folding tended to reduce the effect of the mutations. Greatest rescue was obtained with exposure to sodium 4-phenylbutyrate, a chemical chaperone that increases the expression of heat-shock proteins, and probenecid, a drug used clinically to treat hyperuricemia.


.0007   TUBULOINTERSTITIAL KIDNEY DISEASE, AUTOSOMAL DOMINANT, 1

UMOD, ASN128SER
SNP: rs121917770, ClinVar: RCV002251323

In affected members of a family from Spain (1/96) with autosomal dominant tubulointerstitial kidney disease-1 (ADTKD1; 162000), Turner et al. (2003) identified a heterozygous c.488A-G transition in the UMOD gene that was predicted to result in an asn128-to-ser (N128S) substitution. Functional studies of the variant and studies of patient cells were not performed. The family had previously been reported by Stacey et al. (2003).

In HeLa cells transfected with N128S mutant uromodulin, Williams et al. (2009) found that the mutant protein had delayed maturation (about 25% of wildtype), was retained in the ER, and not trafficked to the plasma membrane, likely due to protein misfolding. The authors suggested that abnormal protein retention in the ER may trigger apoptosis and underlie the mechanism for the pathogenesis of the disorder.


.0008   TUBULOINTERSTITIAL KIDNEY DISEASE, AUTOSOMAL DOMINANT, 1

UMOD, CYS255TYR
SNP: rs121917771, ClinVar: RCV002251324

In affected members of a family from Spain (13/96) with autosomal dominant tubulointerstitial kidney disease-1 (ADTKD1; 162000), Turner et al. (2003) identified a heterozygous c.869G-A transition in the UMOD gene, resulting in a cys255-to-tyr (C255Y) substitution at a conserved residue. Functional studies of the variant and studies of patient cells were not performed. The family had previously been reported by Stacey et al. (2003).

In 14 affected members of a large consanguineous Spanish family with ADTKD1, Rezende-Lima et al. (2004) identified a C255Y mutation in the UMOD gene. These authors stated that the substitution resulted from a c.764G-A transition in exon 4 and that the mutation disrupted a light chain-binding domain. Eleven family members were heterozygous for the mutation, whereas 3 were homozygous. The homozygous individuals had earlier disease onset than the heterozygous individuals, but the report demonstrated that homozygosity for UMOD mutations is not lethal. The clinical phenotype was consistent with medullary cystic kidney disease (MCKD), although the authors noted that cyst formation may be nonspecific secondary effect. This family was also reported as family F4 by Lens et al. (2005).


.0009   TUBULOINTERSTITIAL KIDNEY DISEASE, AUTOSOMAL DOMINANT, 1

UMOD, CYS300GLY
SNP: rs121917772, ClinVar: RCV002251325

In affected members of a family from Spain (20/96) with autosomal dominant tubulointerstitial kidney disease-1 (ADTKD1; 162000), Turner et al. (2003) identified a heterozygous c.1003T-G transversion in the UMOD gene that was predicted to result in a cys300-to-gly (C300G) substitution. Functional studies of the variant and studies of patient cells were not performed. The family had previously been reported by Stacey et al. (2003).

In affected members of a Spanish family (F3) with ADTKD1, Lens et al. (2005) identified heterozygosity for the C300G mutation in exon 5 of the UMOD gene. The phenotype included glomerulocystic kidney disease (GCKD), thus expanding the spectrum of manifestations associated with this mutation.


.0010   TUBULOINTERSTITIAL KIDNEY DISEASE, AUTOSOMAL DOMINANT, 1

UMOD, CYS315ARG
SNP: rs121917773, ClinVar: RCV002251326

In affected members of an Italian family (GCKD1) with autosomal dominant tubulointerstitial kidney disease-1 (ADTKD1; 162000), Rampoldi et al. (2003) identified a c.943T-C transition (c.943T-C, NM_003361) in exon 5 of the UMOD gene, predicted to result in a cys315-to-arg (C315R) substitution. The family had previously been reported by Scolari et al. (1999) with a clinical diagnosis of glomerulocystic kidney disease (GCKD).


.0011   TUBULOINTERSTITIAL KIDNEY DISEASE, AUTOSOMAL DOMINANT, 1

UMOD, VAL273PHE
SNP: rs121917774, ClinVar: RCV001566754, RCV002251327

In 3 Belgian brothers with autosomal dominant tubulointerstitial kidney disease-1 (ADTKD1; 162000), originally reported as family BE2 by Stiburkova et al. (2003), Vylet'al et al. (2006) identified a heterozygous mutation in the UMOD gene resulting in a val273-to-phe (V273F) substitution. Functional studies showed that the V273F-mutant protein was retained in the endoplasmic reticulum and did not reach the plasma membrane.


.0012   TUBULOINTERSTITIAL KIDNEY DISEASE, AUTOSOMAL DOMINANT, 1

UMOD, CYS217GLY
SNP: rs28934583, ClinVar: RCV002251328

In a 36-year-old Belgian woman with autosomal dominant tubulointerstitial kidney disease-1 (ADTKD1; 162000), Dahan et al. (2003) identified a heterozygous c.754T-G transversion in exon 4 of the UMOD gene, resulting in a cys217-to-gly (C217G) substitution. The patient was in end-stage renal failure, and she had a history of gout, with onset at age 21 years. Her father, grandmother, and 1 sister were also affected. Dahan et al. (2003) erroneously reported this mutation as 794T-G, instead of 754T-G, in their paper. In addition, their numbering of the UMOD sequence began at nucleotide 1, rather than at the ATG start codon at nucleotide 106. Following the convention of numbering from the start codon, this mutation should be referred to as c.649T-G (Gross, 2013).

Using transfected MCKD cells, Ma et al. (2012) found that mutant mouse Thp with substitutions corresponding to human cys126 to arg (C126R; 191845.0006) or C217G resulted in protein misfolding, retention of mutant protein in the ER, and reduced surface Thp expression and secretion. In all measures, the C217G substitution within the D8C domain was more severe than the C126R substitution within EGF-like domain-3. Both substitutions also led to reduced expression of the chaperone protein Hsp70 (see HSPA1A; 140550), trapped wildtype Thp in the ER in a dominant-negative manner, and induced apoptosis. Treatments that favored correct protein folding tended to reduce the effect of the mutations. Greatest rescue was obtained with exposure to sodium 4-phenylbutyrate, a chemical chaperone that increases the expression of heat-shock proteins, and probenecid, a drug used clinically to treat hyperuricemia.


.0013   TUBULOINTERSTITIAL KIDNEY DISEASE, AUTOSOMAL DOMINANT, 1

UMOD, CYS248SER
SNP: rs398122388, ClinVar: RCV002251329

In 3 members of a 2-generation family (F6) with autosomal dominant tubulointerstitial kidney disease-1 (ADTKD1; 162000), Zaucke et al. (2010) identified a heterozygous c.743G-C transversion in the UMOD gene, resulting in a cys248-to-ser substitution (C248S) at a conserved residue in the D8C domain. The father exhibited hypertension, gout, congestive heart failure, varices, and an aortic aneurysm. Renal biopsy showed tubular atrophy, tubulointerstitial fibrosis, and a thickened tubular basement membrane.


.0014   TUBULOINTERSTITIAL KIDNEY DISEASE, AUTOSOMAL DOMINANT, 1

UMOD, GLN316PRO
SNP: rs1555487318, ClinVar: RCV002251372

In affected members of 2 apparently unrelated Spanish families (F1 and F2) with autosomal dominant tubulointerstitial kidney disease-1 (ADTKD1; 162000), Lens et al. (2005) identified a heterozygous c.947A-C transversion in exon 5 of the UMOD gene, resulting in a gln316-to-pro (Q316P) substitution at a conserved residue. The mutation segregated with the phenotype in the families and was not found in 100 control chromosomes. The clinical phenotype in family 1 was consistent with medullary cystic kidney disease (MCKD), whereas the clinical phenotype in family 2 was consistent with familial juvenile hyperuricemic nephropathy (FJHN). The results demonstrated that these variable clinical manifestations are part of the same disease entity, referred to as ADTKD1. Although functional studies of the variant were not performed, the authors suggested that the mutation could alter the tertiary structure of UMOD, leading to defects in vesicular transport in the thick ascending loop of Henle.


.0015   TUBULOINTERSTITIAL KIDNEY DISEASE, AUTOSOMAL DOMINANT, 1

UMOD, CYS148TRP
SNP: rs1965741911, ClinVar: RCV002251407

In affected members of an Italian family (MCKD9) with autosomal dominant tubulointerstitial kidney disease-1 (ADTKD1; 162000), Rampoldi et al. (2003) identified a heterozygous c.444T-G transversion (c.444T-G, NM_003361) in exon 4 of the UMOD gene, resulting in a cys148-to-trp (C148W) substitution at a conserved residue.

Bernascone et al. (2010) generated a transgenic mouse model with a Umod C147W mutation corresponding to the human UMOD C148W mutation. The mutant mice showed tubulointerstitial fibrosis with inflammatory cell infiltration, tubule dilation, and selective damage to the epithelial cells lining the thick ascending limb of the loop of Henle (TAL), leading to mild renal failure. Umod was retained in the ER of expressing cells, leading to ER hyperplasia. The authors suggested that impaired TAL function is a consequence of a gain-of-function effect of UMOD mutations.


.0016   TUBULOINTERSTITIAL KIDNEY DISEASE, AUTOSOMAL DOMINANT, 1

UMOD, ASP196ASN
SNP: rs1965729129, ClinVar: RCV002251408

In 3 affected members of a family (family 5) with autosomal dominant tubulointerstitial kidney disease-1 (ADTKD1; 162000), Williams et al. (2009) identified a heterozygous c.586G-A transition in exon 3 of the UMOD gene, resulting in an asp196-to-asn (D196N) substitution in the cysteine-rich domain. In vitro functional expression studies in HeLa cells transfected with the mutation showed that it caused a 50% delay in maturation compared to wildtype. The mutant protein was retained in the ER and showed decreased expression at the plasma membrane. The findings suggested that abnormal protein retention in the ER may trigger apoptosis and underlie the mechanism for disease pathogenesis.


REFERENCES

  1. Bachmann, S., Mutig, K., Bates, J., Welker, P., Geist, B., Gross, V., Luft, F. C., Alenina, N., Bader, M., Thiele, B. J., Prasadan, K., Raffi, H. S., Kumar, S. Renal effects of Tamm-Horsfall protein (uromodulin) deficiency in mice. Am. J. Physiol. Renal Physiol. 288: F559-F567, 2005. [PubMed: 15522986] [Full Text: https://doi.org/10.1152/ajprenal.00143.2004]

  2. Bernascone, I., Janas, S., Ikehata, M., Trudu, M., Corbelli, A., Schaeffer, C., Rastaldi, M. P., Devuyst, O., Rampoldi, L. A transgenic mouse model for uromodulin-associated kidney diseases shows specific tubulo-interstitial damage, urinary concentrating defect and renal failure. Hum. Molec. Genet. 19: 2998-3010, 2010. [PubMed: 20472742] [Full Text: https://doi.org/10.1093/hmg/ddq205]

  3. Dahan, K., Devuyst, O., Smaers, M., Vertommen, D., Loute, G., Poux, J. M., Viron, B., Jacquot, C., Gagnadoux, M. F., Chauveau, D., Buchler, M., Cochat, P., Cosyns, J. P., Mougenot, B., Rider, M. H., Antignac, C., Verellen-Dumoulin, C., Pirson, Y. A cluster of mutations in the UMOD gene causes familial juvenile hyperuricemic nephropathy with abnormal expression of uromodulin. J. Am. Soc. Nephrol. 14: 2883-2893, 2003. [PubMed: 14569098] [Full Text: https://doi.org/10.1097/01.asn.0000092147.83480.b5]

  4. Deng, Z., Johnson, K., Engleward, B. P., Lane, S., Callen, D. F., Samson, L. D., Davisson, M. T., Siciliano, M. J. New regions of conserved synteny and linkage between human chromosome 16p12-p13 and mouse chromosomes 16 and 11. (Abstract) Cytogenet. Cell Genet. 68: 180 only, 1995.

  5. Fairbanks, L. D., Cameron, J. S., Venkat-Raman, G., Rigden, S. P. A., Rees, L., Van't Hoff, W., Mansell, M., Pattison, J., Goldsmith, D. J. A., Simmonds, H. A. Early treatment with allopurinol in familial juvenile hyperuricaemic nephropathy (FJHN) ameliorates the long-term progression of renal disease. Quart. J. Med. 95: 597-607, 2002. [PubMed: 12205338] [Full Text: https://doi.org/10.1093/qjmed/95.9.597]

  6. Fukuoka, S.-I., Matsuda, Y. Assignment of the Tamm-Horsfall protein/uromodulin gene (Umod) to mouse chromosome bands 7F1-F2 and rat chromosome bands 1q36-q37 by in situ hybridization. Cytogenet. Cell Genet. 79: 241-242, 1997. [PubMed: 9605864] [Full Text: https://doi.org/10.1159/000134735]

  7. Greimel, P., Jabs, S., Storch, S., Cherif, S., Honke, K., Braulke, T., Thiem, J. In vitro sulfation of N-acetyllactosaminide by soluble recombinant human beta-Gal-3-prime-sulfotransferase. Carbohydr. Res. 341: 918-924, 2006. [PubMed: 16516177] [Full Text: https://doi.org/10.1016/j.carres.2006.02.006]

  8. Gross, M. B. Personal Communication. Baltimore, Md. 2/11/2013.

  9. Hart, T. C., Gorry, M. C., Hart, P. S., Woodard, A. S., Shihabi, Z., Sandhu, J., Shirts, B., Xu, L., Zhu, H., Barmada, M. M., Bleyer, A. J. Mutations of the UMOD gene are responsible for medullary cystic kidney disease 2 and familial juvenile hyperuricaemic nephropathy. J. Med. Genet. 39: 882-892, 2002. [PubMed: 12471200] [Full Text: https://doi.org/10.1136/jmg.39.12.882]

  10. Jeanpierre, C., Whitmore, S. A., Austruy, E., Cohen-Salmon, M., Callen, D. F., Junien, C. Chromosomal assignment of the uromodulin gene (UMOD) to 16p13.11. Cytogenet. Cell Genet. 62: 185-187, 1993. [PubMed: 8382593] [Full Text: https://doi.org/10.1159/000133470]

  11. Lens, X. M., Banet, J. F., Outeda, P., Barrio-Lucia, V. A novel pattern of mutation in uromodulin disorders: autosomal dominant medullary cystic kidney disease type 2, familial juvenile hyperuricemic nephropathy, and autosomal dominant glomerulocystic kidney disease. Am. J. Kidney Dis. 46: 52-57, 2005. [PubMed: 15983957] [Full Text: https://doi.org/10.1053/j.ajkd.2005.04.003]

  12. Ma, L., Liu, Y., El-Achkar, T. M., Wu, X.-R. Molecular and cellular effects of Tamm-Horsfall protein mutations and their rescue by chemical chaperones. J. Biol. Chem. 287: 1290-1305, 2012. [PubMed: 22117067] [Full Text: https://doi.org/10.1074/jbc.M111.283036]

  13. Massari, P. U., Hsu, C. H., Barnes, R. V., Fox, I. H., Gikas, P. W., Weller, J. M. Familial hyperuricemia and renal disease. Arch. Intern. Med. 140: 680-684, 1980. [PubMed: 7396593]

  14. McBride, M. B., Ridgen, S., Haycock, G. B., Dalton, N., Van't Hoff, W., Rees, L., Venkat-Raman, G. V., Moro, F., Ogg, C. S., Cameron, J. S., Simmonds, H. A. Presymptomatic detection of familial juvenile hyperuricaemic nephropathy in children. Pediat. Nephrol. 12: 357-364, 1998. [PubMed: 9686952] [Full Text: https://doi.org/10.1007/s004670050466]

  15. McKenzie, J. K., McQueen, E. G. Immunofluorescent localization of Tamm-Horsfall mucoprotein in human kidney. J. Clin. Path. 22: 334-339, 1969. [PubMed: 4891482] [Full Text: https://doi.org/10.1136/jcp.22.3.334]

  16. Mo, L., Huang, H.-Y., Zhu, X.-H., Shapiro, E., Hasty, D. L., Wu, X.-R. Tamm-Horsfall protein is a critical renal defense factor protecting against calcium oxalate crystal formation. Kidney Int. 66: 1159-1166, 2004. [PubMed: 15327412] [Full Text: https://doi.org/10.1111/j.1523-1755.2004.00867.x]

  17. Mo, L., Zhu, X.-H., Huang, H.-Y., Shapiro, E., Hasty, D. L., Wu, X.-R. Ablation of the Tamm-Horsfall protein gene increases susceptibility of mice to bladder colonization by type 1-fimbriated Escherichia coli. Am. J. Physiol. Renal Physiol. 286: F795-F802, 2004. [PubMed: 14665435] [Full Text: https://doi.org/10.1152/ajprenal.00357.2003]

  18. Muchmore, A. V., Decker, J. M. Uromodulin: a unique 85-kilodalton immunosuppressive glycoprotein isolated from urine of pregnant women. Science 229: 479-481, 1985. [PubMed: 2409603] [Full Text: https://doi.org/10.1126/science.2409603]

  19. Pennica, D., Kohr, W. J., Kuang, W.-J., Glaister, D., Aggarwal, B. B., Chen, E. Y., Goeddel, D. V. Identification of human uromodulin as the Tamm-Horsfall urinary glycoprotein. Science 236: 83-88, 1987. [PubMed: 3453112] [Full Text: https://doi.org/10.1126/science.3453112]

  20. Pook, M. A., Jeremiah, S., Scheinman, S. J., Povey, S., Thakker, R. V. Localization of the Tamm-Horsfall glycoprotein (uromodulin) gene to chromosome 16p12.3-16p13.11. Ann. Hum. Genet. 57: 285-290, 1993. [PubMed: 8179291] [Full Text: https://doi.org/10.1111/j.1469-1809.1993.tb00902.x]

  21. Rampoldi, L., Caridi, G., Santon, D., Boaretto, F., Bernascone, I., Lamorte, G., Tardanico, R., Dagnino, M., Colussi, G., Scolari, F., Ghiggeri, G. M., Amoroso, A., Casari, G. Allelism of MCKD, FJHN and GCKD caused by impairment of uromodulin export dynamics. Hum. Molec. Genet. 12: 3369-3384, 2003. [PubMed: 14570709] [Full Text: https://doi.org/10.1093/hmg/ddg353]

  22. Rezende-Lima, W., Parreira, K. S., Garcia-Gonzalez, M., Riveira, E., Banet, J. F., Lens, X. M. Homozygosity for uromodulin disorders: FJHN and MCKD-type 2. Kidney Int. 66: 558-563, 2004. [PubMed: 15253706] [Full Text: https://doi.org/10.1111/j.1523-1755.2004.00774.x]

  23. Schenk, E. A., Schwartz, R. H., Lewis, R. A. Tamm-Horsfall mucoprotein. I. Localization in the kidney. Lab Invest. 25: 92-95, 1971. [PubMed: 4933557]

  24. Scolari, F., Puzzer, D., Amoroso, A., Caridi, G., Ghiggeri, G. M., Maiorca, R., Aridon, P., De Fusco, M., Ballabio, A., Casari, G. Identification of a new locus for medullary cystic disease, on chromosome 16p12. Am. J. Hum. Genet. 64: 1655-1660, 1999. [PubMed: 10330352] [Full Text: https://doi.org/10.1086/302414]

  25. Stacey, J. M., Turner, J. J. O., Harding, B., Nesbit, M. A., Kotanko, P., Lhotta, K., Puig, J. G., Torres, R. J., Thakker, R. V. Genetic mapping studies of familial juvenile hyperuricemic nephropathy on chromosome 16p11-p13. J. Clin. Endocr. Metab. 88: 464-470, 2003. [PubMed: 12519891] [Full Text: https://doi.org/10.1210/jc.2002-021268]

  26. Stiburkova, B., Majewski, J., Hodanova, K., Ondrova, L., Jerabkova, M., Zikanova, M., Vylet'al, P., Sebesta, I., Marinaki, A., Simmonds, A., Matthijs, G., Fryns, J.-P., Torres, R., Puig, J. G., Ott, J., Kmoch, S. Familial juvenile hyperuricaemic nephropathy (FJHN): linkage analysis in 15 families, physical and transcriptional characterisation of the FJHN critical region on chromosome 16p11.2 and the analysis of seven candidate genes. Europ. J. Hum. Genet. 11: 145-154, 2003. [PubMed: 12634862] [Full Text: https://doi.org/10.1038/sj.ejhg.5200937]

  27. Stiburkova, B., Majewski, J., Sebesta, I., Zhang, W., Ott, J., Kmoch, S. Familial juvenile hyperuricemic nephropathy: localization of the gene on chromosome 16p11.2--and evidence for genetic heterogeneity. Am. J. Hum. Genet. 66: 1989-1994, 2000. [PubMed: 10780922] [Full Text: https://doi.org/10.1086/302936]

  28. Tamm, I., Horsfall, F. L., Jr. A mucoprotein derived from human urine which reacts with influenza, mumps, and Newcastle disease viruses. J. Exp. Med. 95: 71-97, 1952. [PubMed: 14907962] [Full Text: https://doi.org/10.1084/jem.95.1.71]

  29. Turner, J. J. O., Stacey, J. M., Harding, B., Kotanko, P., Lhotta, K., Puig, J. G., Roberts, I., Torres, R. J., Thakker, R. V. UROMODULIN mutations cause familial juvenile hyperuricemic nephropathy. J. Clin. Endocr. Metab. 88: 1398-1401, 2003. [PubMed: 12629136] [Full Text: https://doi.org/10.1210/jc.2002-021973]

  30. Vylet'al, P., Kublova, M., Kalbacova, M., Hodanova, K., Baresova, V., Stiburkova, B., Sikora, J., Hulkova, H., Zivny, J., Majewski, J., Simmonds, A., Fryns, J.-P., Venkat-Raman, G., Elleder, M., Kmoch, S. Alterations of uromodulin biology: a common denominator of the genetically heterogeneous FJHN/MCKD syndrome. Kidney Int. 70: 1155-1169, 2006. [PubMed: 16883323] [Full Text: https://doi.org/10.1038/sj.ki.5001728]

  31. Weiss, G. L., Stanisich, J. J., Sauer, M. M., Lin, C.-W., Eras, J., Zyla, D. S., Truck, J., Devuyst, O., Aebi, M., Pilhofer, M., Glockshuber, R. Architecture and function of human uromodulin filaments in urinary tract infections. Science 369: 1005-1010, 2020. [PubMed: 32616672] [Full Text: https://doi.org/10.1126/science.aaz9866]

  32. Williams, S. E., Reed, A. A. C., Galvanovskis, J., Antignac, C., Goodship, T., Karet, F. E., Kotanko, P., Lhotta, K., Moriniere, V., Williams, P., Wong, W., Rorsman, P., Thakker, R. V. Uromodulin mutations causing familial juvenile hyperuricaemic nephropathy lead to protein maturation defects and retention in the endoplasmic reticulum. Hum. Molec. Genet. 18: 2963-2974, 2009. [PubMed: 19465746] [Full Text: https://doi.org/10.1093/hmg/ddp235]

  33. Yang, H., Wu, C., Zhao, S., Guo, J. Identification and characterization of D8C, a novel domain present in liver-specific LZP, uromodulin and glycoprotein 2, mutated in familial juvenile hyperuricaemic nephropathy. FEBS Lett. 578: 236-238, 2004. [PubMed: 15589826] [Full Text: https://doi.org/10.1016/j.febslet.2004.10.092]

  34. Zaucke, F., Boehnlein, J. M., Steffens, S., Polishchuk, R. S., Rampoldi, L., Fischer, A., Pasch, A., Boehm, C. W. A., Baasner, A., Attanasio, M., Hoppe, B., Hopfer, H., Beck, B. B., Sayer, J. A., Hildebrandt, F., Wolf, M. T. F. Uromodulin is expressed in renal primary cilia and UMOD mutations result in decreased ciliary uromodulin expression. Hum. Molec. Genet. 19: 1985-1997, 2010. [PubMed: 20172860] [Full Text: https://doi.org/10.1093/hmg/ddq077]


Contributors:
Cassandra L. Kniffin - updated : 01/26/2021
Ada Hamosh - updated : 12/04/2020
Carol A. Bocchini - updated : 10/18/2017
George E. Tiller - updated : 9/9/2013
Cassandra L. Kniffin - updated : 2/13/2013
Matthew B. Gross - updated : 2/11/2013
Patricia A. Hartz - updated : 2/8/2013
George E. Tiller - updated : 6/25/2010
Marla J. F. O'Neill - updated : 10/15/2009
George E. Tiller - updated : 1/10/2006
Patricia A. Hartz - updated : 4/19/2005
John A. Phillips, III - updated : 3/29/2004
Victor A. McKusick - updated : 6/30/2003
Victor A. McKusick - updated : 5/28/1998

Creation Date:
Victor A. McKusick : 6/6/1994

Edit History:
alopez : 02/09/2021
ckniffin : 01/26/2021
mgross : 12/04/2020
carol : 10/19/2017
carol : 10/18/2017
alopez : 09/09/2013
alopez : 9/9/2013
tpirozzi : 9/5/2013
tpirozzi : 9/5/2013
carol : 3/4/2013
ckniffin : 2/13/2013
mgross : 2/11/2013
mgross : 2/11/2013
terry : 2/8/2013
mgross : 8/3/2011
carol : 12/10/2010
wwang : 7/15/2010
terry : 6/25/2010
alopez : 1/6/2010
terry : 10/21/2009
wwang : 10/16/2009
terry : 10/15/2009
carol : 6/18/2008
terry : 12/17/2007
wwang : 2/10/2006
terry : 1/10/2006
mgross : 4/25/2005
terry : 4/19/2005
alopez : 3/29/2004
tkritzer : 7/15/2003
tkritzer : 7/7/2003
terry : 6/30/2003
terry : 6/1/1998
terry : 5/28/1998
jason : 6/6/1994