Skip to content
BY-NC-ND 3.0 license Open Access Published by De Gruyter February 12, 2014

Transthyretin: a multifaceted protein

  • Marta Vieira

    Marta Vieira (left) has a degree in Biochemistry from the University of Porto and obtained in 2013 her PhD in Biomedical Sciences at the same University participating in several research projects at the Molecular Neurobiology group at the Institute for Cellular and Molecular Biology at Porto University.

    and Maria João Saraiva

    Maria João Saraiva (right) received a BSc in Biology from the University of Porto in 1976, and an MSc in Biochemistry from the University of London, in 1978. Between 1980 and 1984, she studied for a PhD in biochemistry at the University of Porto, and qualified as Professor of Biochemistry in 1991. She worked for different periods as a Visiting Scientist at the College of Physicians and Surgeons at Columbia University, New York. Maria J. Saraiva currently holds a number of positions at the Universidade do Porto, Portugal: she is Vice Director at the Institute for Molecular and Cellular Biology (IBMC); a Professor of Biochemistryat the Biomedical Institute; and Director of the Molecular Neurobiology Group at the IBMC.

    EMAIL logo
From the journal BioMolecular Concepts

Abstract

Transthyretin is a highly conserved homotetrameric protein, mainly synthetized by the liver and the choroid plexus of brain. The carrier role of TTR is well-known; however, many other functions have emerged, namely in the nervous system. Behavior, cognition, neuropeptide amidation, neurogenesis, nerve regeneration, axonal growth and 14-3-3ζ metabolism are some of the processes where TTR has an important role. TTR aggregates are responsible for many amyloidosis such as familial amyloidotic polyneuropathy and cardiomyopathy. Normal TTR can also aggregate and deposit in the heart of old people and in preeclampsia placental tissue. Differences in TTR levels have been found in several neuropathologies, but its neuroprotective role, until now, was described in ischemia and Alzheimer’s disease. The aim of this review is to stress the relevance of TTR, besides its well-known role on transport of thyroxine and retinol-binding protein.

Introduction

In 1942, a protein migrating ahead of albumin during electrophoresis of human plasma (1) and cerebrospinal fluid (CSF) (2) samples, was identified and denominated ‘prealbumin’. The discovery that prealbumin could bind thyroid hormones (THs) led to a change of its name to ‘thyroxine-binding prealbumin’ (TBPA) (3). Later, it was found that TBPA could also bind retinol-binding protein (RBP) (4). In 1981, the International Union of Biochemists adopted the name ‘transthyretin’ (TTR) (5) because of its well-known role in the transport of the THs thyroxine and retinol (vitamin A) through binding to RBP.

TTR is a highly conserved protein that is found in several vertebrate species (6, 7). Recently, sequences homologous to TTR, known as transthyretin-like proteins (TLPs), have been found in bacteria, nematodes and plants. In Escherichia coli and Caenorhabditis elegans TLPs form homotetramers, like TTR, although without the ability to bind T4 (8).

Structure

The first X-ray crystal structure of TTR was reported in 1971 (9). TTR is a 54,980 Daltons (Da) homotetrameric protein; each subunit has 13,745 Da and is composed of 127 amino acids (10). Native TTR has a globular shape with an overall size of 70 Å×55 Å×50 Å and a central hydrophobic channel. Each monomer consists of eight antiparallel β-strands (A through H), which are organized into two four-stranded β-sheets (DAGH and CBEF) and only a short α-helix located on β-strand E (11). A dimer is formed when β-strands F and H of each subunit interact by hydrogen bonds. Tetramer formation results from interaction of the residues of the loops that join β-strands G to H and A to B. It is known that low pH promotes the dissociation of TTR tetramer into monomers. The neutral crystal structure of TTR demonstrated double protonation of His88 to break the hydrogen-bond network, causing destabilization of the TTR tetramer (12).

Expression

TTR is mainly synthetized by the liver (13) and the choroid plexus of brain (14), which are the sources of TTR in plasma and CSF, respectively. In humans, 90% of plasma TTR is secreted from liver and its concentration ranges from 20 to 40 mg/dl (15). TTR levels in plasma change with age: they are decreased in healthy newborns when compared with adults (16, 17) and start to decline after 50 years of age (18).

Synthesis of TTR by epithelial cells of the choroid plexus is the main source of CSF TTR (14). TTR concentration in CSF ranges from 5 to 20 mg/l (19), representing approximately 25% of the total CSF protein content (20). The choroid plexus has 11 times more mRNA than the liver, normalized for tissue weight, and synthetizes TTR 13 times faster than the liver (21).

TTR synthesis in brain areas, other than the choroid plexus, has been a controversial subject. The presence of TTR mRNA in murine or human brains has been detected in brain areas such as the cortex, hippocampus or cerebellum (22–25). Some authors claimed that brain TTR may be caused by neuronal synthesis of the protein in these tissues; while others proved, by laser microdissection technology and in situ hybridization, that TTR is not produced in brain parenchyma, suggesting that TTR contamination by choroid plexus may induce false-positive results concerning sites of TTR synthesis (26).

Besides the liver and the choroid plexus, TTR synthesis has been described in several other tissues. TTR is highly transcribed and translated in the retinal pigment epithelium (RPE), a monolayer of cells that acts as blood barrier for the retina (27). TTR is also produced in the pancreatic islet of Langerhans (α cells) (28, 29), and to a small extent in the heart, skeletal muscle, spleen (30), visceral yolk sac endoderm (31), pineal gland (32) and trophoblasts of human placenta (33). Although the expression of TTR in the stomach was previously described (30), a specific gastric cellular population of ghrelin cells was recently identified as TTR producers (34).

Metabolism

Although TTR production has been extensively studied, its catabolism is not fully understood. The biological half-life of TTR is about 2–3 days in humans (35), 23h in monkeys (36) and 29h in Buffalo rats (37). In 1988, Makover and colleagues demonstrated that the major sites of TTR degradation were the liver, muscle and skin. In their studies, 36–38% of total body TTR degradation occurred in liver, 12–15% in muscle and 8–10% in skin. In kidneys, adipose tissue, testis and gastrointestinal tract (GI), the degradation rate of body TTR was 1–8%, whereas <1% was degraded by the other tissues examined (38). No evidence was found of TTR degradation in tissues of the nervous system. Liver and kidney were the most active organs of TTR catabolism, per gram of wet weight. TTR internalization in liver and kidney is receptor-mediated. Renal uptake of TTR was shown to be megalin-[also known as low-density lipoprotein-related protein 2 (LRP2)] mediated (39). Megalin is an endocytic multi-ligand receptor of the low-density lipoprotein (LDL) receptor family, that is expressed in the epithelium of renal proximal tubes, among other epithelia. Sousa et al. demonstrated that TTR uptake in liver was mediated by a receptor member of LDL family sensitive to receptor-associated protein (RAP) (40). Inhibition of TTR uptake by RAP suggested a common pathway between TTR and lipoproteins metabolism. As megalin is not expressed in liver, further studies need to be performed in order to clarify TTR internalization in this organ.

TTR synthesis and secretion by placental trophoblasts was described (33, 41). Secretion of TTR from trophoblasts to maternal placental circulation is followed by trophoblasts uptake of TTR-T4 complex to the fetal circulation (42) and is important for fetal development. This uptake was suggested to occur through a LRP receptor.

Functions

TTR has been mainly recognized by its role as a carrier protein of THs and retinol in plasma and CSF. However, during the last years, a role in proteolysis and in several functions in the nervous system has also been proposed.

Transport of T4

Thyroid hormones are iodinated compounds essential for development, tissue differentiation and regulation of metabolic balance in mammals. In recent years a role in cell migration, signaling, myelination and promotion of neurite outgrowth has also been proposed (43, 44).

The thyroid gland synthetizes three THs: tetraiodothyroxine (T4), triiodothyronine (T3) and a biologically inactive reverse T3 (rT3).

T4 is the most abundant TH secreted by the thyroid gland and circulates in plasma bound to thyroid hormone-binding proteins (THBPs) thyroxin-binding globulin (TBG), TTR and albumin. TBG has the strongest affinity to T4, carrying 65% of the hormone (45); 15% of human plasma T4 is transported by TTR and about 10% by albumin; finally, 0.03–0.05% circulates unbound or in a free form (46). In rodents, 50% of total T4 is carried by TTR (47). In CSF of both rodents and humans, TTR is the main carrier of T4, transporting 80% of the hormone (48). The homotetrameric structure of native TTR forms a central hydrophobic channel with two binding sites for T4 (49). As these binding sites exhibit negative cooperativity, just one molecule of T4 is transported by TTR (50).

The delivery of T4 in cells is not a consensual subject; while some defend that uptake of T4 occurs bound to the carrier proteins, others claim that T4 enters the cell by passive diffusion after dissociation from the carrier protein (51, 52). Studies on TTR null mice (53) support the free T4 tissue uptake hypothesis. TTR null mice exhibited 50% reduction of total T4 in blood and 30% in CSF, when compared with wild-type animals, whereas the levels of free T4 and total circulating T3 were unaltered. Several parameters measured in this strain of mice indicated the animals to be euthyroid (54). Taken together, the results suggest that TTR is not pivotal to TH metabolism, even in conditions of increased hormone demand as cold exposition or thyroidectomy (55).

The redundant role of TTR was also described for other THBPs, such as albumin in rats (56) and TBG in humans (57, 58). A critical role for TTR on T4 transport across the placenta and delivery to the fetus was recently described (42, 59). Further studies need to be performed to clarify the role of TTR on T4 delivery into tissues.

Transport of retinol

Retinol, or vitamin A, and related metabolites are obtained from the diet. Oxidation of retinol originates retinoic acid, which is very important in several functions including vision, reproduction, growth and development (60). It also modulates neurogenesis, neuronal survival and synaptic plasticity in hippocampus, olfactory bulb and hypothalamus (61). Retinoids can regulate cell differentiation, neurite outgrowth and protection against oxidative stress (62). Several studies suggest vitamin A as a possible therapeutic agent in Alzheimer disease (AD): (i) the disruption of retinoid signaling induced deposition of Aβ in the cerebral blood vessels (63); (ii) vitamin A inhibited amyloid fibril formation (64) and (iii) in vivo studies revealed that administration of vitamin A decreased Aβ deposition and improved spatial learning and memory (65).

The transport of retinol in circulation occurs through RBP (66). This molecule is mainly synthetized in the liver and secreted to plasma after retinol binding. TTR associates to the RBP-retinol complex before secretion into the plasma (4). The TTR-RBP complex is a very stable form of retinol transport, allowing its delivery to cells and is important to prevent RBP from being filtered and degraded in the kidney (67, 68).

TTR tetramer has four RBP-binding sites, two in each dimer at the protein’s surface; because of steric hindrance, just two RBP are transported by each TTR molecule. Under physiological conditions, due to low RBP levels comparatively with TTR, just one molecule of RBP is transported by the TTR tetramer (69, 70). T4 binding to TTR is not influenced by RBP binding (4).

Studies in TTR null mice showed a dramatic reduction of retinol and RBP plasma levels (around 95%) when compared with wild-type littermates. This finding could be explained by increased renal filtration of the retinol-RBP complex (70). Increased hepatic RBP levels were found in TTR null mice (71). However, in vitro and in vivo studies demonstrated that RBP liver secretion from plasma was unchanged (70), which indicates that diminished levels of RBP and retinol in plasma are not caused by a failure in secretion.

Symptoms of vitamin A deficiency, such as weight loss, infections and eye abnormalities were not observed in TTR mutant mice (70), suggesting that TTR role on RBP-retinol transport, does not play a critical role on retinol metabolism.

Retinol uptake was suggested to be mediated by a TTR-independent membrane receptor (72). Stra6 (a multi-transmembrane domain protein) was reported as a mediator of RBP4 binding to cell membranes and as crucial for cellular uptake of retinol (73). Moreover, a new retinol transporter has recently been identified, RBPR2, and suggested to play a role in retinol absorption (74).

Proteolytic activity

Another important function of TTR besides its role in transport of T4 and retinol is its proteolytic activity on several substrates. A small fraction of plasma TTR (1–2%) is carried by high-density lipoproteins (HDL) through binding to apolipoprotein (apo) A-I (75). The TTR-apoA-I interaction was further investigated and TTR was described as a non-canonical serine protease capable to cleave apoA-I carboxyl terminal domain (76) and to reduce cholesterol efflux (77). TTR has also the ability to cleave neuropeptide Y (NPY) (78) and Aβ peptide (79). Cleavage of Aβ can occur at several different sites, and the resulting peptides were shown to have decreased amyloidogenic potential when compared with the complete peptide. TTR was also able to degrade aggregated forms of Aβ; inhibition of TTR activity resulted in increased Aβ fibril formation (79). TTR proteolytic role on NPY and Aβ peptide need further investigation to determine its functional role in the nervous system.

TTR in the nervous system

Several studies have shown different roles for TTR in nervous system physiology. Behavior, cognition, neuropeptide amidation, neurogenesis, nerve regeneration, axonal growth and 14-3-3ζ metabolism are some of the known processes influenced by TTR as studied in TTR null mice.

TTR null animals are viable, phenotypically similar to wild-type and heterozygous littermates, and fertile (53). This strain presents reduced signs of depressive-like behavior, increased exploratory activity and anxiety (80). Increased levels of norepinephrine in the limbic forebrain could be a possible explanation for the observed phenotype. TTR null mice presented increased levels of NPY in dorsal root ganglia (DRG), sciatic nerve, spinal cord, hippocampus, cortex and CSF ascribed to up-regulation of peptidylglycine α-amidating monooxygenase (PAM) – the only enzyme that amidates neuropeptides, being crucial for the maturation process of NPY (81, 82). These findings corroborate the importance of TTR in modulating depressive behavior.

Cognitive performance analysis of young/adult TTR null mice showed memory impairment when compared with wild-type littermates (24, 83, 84). With aging, TTR wild-type animals presented worsened cognitive performance, attributable to reduced levels of CSF TTR, a fact that enhances the important role of TTR in cognition (83).

Increased locomotor activity in young/adult TTR null animals was confirmed by Fleming et al.; in older mice, a sensorimotor impairment was observed (85). No morphological differences in sciatic nerves and cerebellum were found in TTR null animals that could explain the absence of sensorimotor impairment at young ages. However, under nerve crush conditions, absence of TTR slowed nerve regeneration (85). TTR null mice have slower recovery of locomotor activity and slower nerve conduction velocity. Neuropathological parameters such as decreased levels of myelinated and unmyelinated axons were also observed in TTR null animals when compared with wild-type littermates. TTR properties as a nerve regeneration enhancer were further demonstrated when TTR delivery to crushed sciatic nerves rescued the regeneration phenotype of TTR null animals (86).

TTR has also the capacity of inducing neurite outgrowth in DGR and PC12 cells, independently of its ligands (85). Neuritogenic activity of TTR in DRG neurons depends on its internalization, a process that is clathrin-dependent and megalin-mediated. In vivo studies in a mouse model with reduced levels of megalin, demonstrated decreased nerve regeneration capacity. Thus, reduced megalin levels impair TTR action as an enhancer of regeneration (86).

The binding of TTR to glucose-regulated protein (Grp)78 at the plasma membrane of β-cells and its internalization was recently demonstrated (87) and hypothesized to be essential for insulin release and protection against β-cells apoptosis (88).

Finally a role for TTR in 14-3-3ζ metabolism has been suggested. Hippocampus of young TTR null mice presented lower levels of 14-3-3ζ protein, but no changes in gene expression. Lysosomal degradation of 14-3-3ζ in the absence of TTR was the mechanism proposed for the reduced 14-3-3ζ levels (89). Hippocampal slice cultures of TTR null mice did not present increased cellular death when compared wild-type animals (90), suggesting that the decreased levels of 14-3-3ζ observed in these animals are not associated with increased cellular death. Furthermore, in Creutzfeldt-Jakob disease patients, 14-3-3 proteins are used as a surrogate marker for the in vivo diagnosis of the disease, whereas TTR levels are the same as compared with healthy controls (91, 92). Taken together, these results suggest that the regulation of 14-3-3ζ levels by TTR is not associated with cellular death. Further studies should be performed to dissect the mechanism and the consequences of this regulation.

TTR in diseases of the nervous system

Differences in TTR levels and/or increased oxidation are found in several neuropathologies, such as Guillain-Barré syndrome (GBS), frontotemporal dementia (FDT), amyotrophic lateral sclerosis (ALS) and Parkinson’s disease (PD). In GBS and FDT, TTR levels are elevated as compared to controls, in plasma and CSF (93, 94), whereas in ALS TTR was up regulated in rapid progression of the disease when compared with slow progression of the disease (95). Finally, plasma and CSF from PD patients present increased TTR levels when compared to controls, which led to suggest TTR as an early marker of the disease (96). Further studies should be performed in order to dissect the underlying mechanisms behind TTR role in PD.

The neuroprotective role of TTR in the central nervous system has been described in more detail in ischemia and AD and will be discussed in the next sections.

Cerebral ischemia

Ischemia is a significant cause of brain injury worldwide, leading to high mortality, physical and cognitive incapacities. TTR was identified as being differentially expressed in fluids or tissues after ischemia (97–99) and a mice model of permanent middle cerebral artery occlusion (pMCAO) showed TTR as a neuroprotective molecule. TTR null mice heterozygous for heatshock transcription factor 1 (HSF1) (TTR-/-/HSF1/+/-) presented increased infarct area, cerebral edema and microglial-leukocyte response when compared with TTR wild-type animals (TTR+/+/HSF+/-), 24h after pMCAO. Interestingly, TTR was localized in disintegrated β-tubulin III-positive neurons and cell debris. Elimination of TTR secreted by liver after treatment with RNAi had no effect on the distribution of TTR in endangered neurons, indicating that TTR mobilization to neurons was of CSF origin. These results suggested the neuroprotective role of CSF TTR in cerebral ischemia (100).

An epidemiologic study revealed that the population heterozygous for T119M (a TTR mutant with increased tetrameric stability) have decreased risk of cerebrovascular disease and increased life expectancy when compared with noncarriers individuals (101).

The mechanisms behind these observations warrant further investigation.

Alzheimer’s disease

AD is the most common form of dementia, affecting millions of people worldwide. The two main histopathological marks of AD are neurofibrillary tangles (aggregates of hyperphosphorylated tau protein) and senile plaques (aggregates of Aβ peptide) (102).

A role for TTR in AD has been suggested by several groups. The first description of decreased TTR levels in CSF of AD patients dates back to 1986 (103). Recently, with the use of powerful tools, a 2-fold decrease in TTR levels in CSF of AD patients was demonstrated (92, 104). As observed in CSF, TTR plasma levels are decreased in AD patients and mild-cognitive impairment (MCI) when compared with non-demented controls, suggesting TTR as an early AD biomarker (105, 106).

TTR is able to bind Aβ peptides, preventing the formation of amyloid fibrils (107). Analysis of Aβ aggregation kinetics, showed that, in the presence of TTR, the aggregation rate of this peptide is decreased (108).

Animal models of AD are a useful tool to dissect the mechanism behind the disease. Different animal models have been used but TTR role in AD is controversial.

Initially, a protective role of TTR was described in APPswe/PS1ΔE9 animals. In this strain of mice [double transgenic mice expressing the human amyloid precursor (APP) with a Swedish mutation and a human presenilin (PS) deletion] Aβ and amyloid deposits were decreased when the animals were exposed to an enriched environment (large cages, running wheels, colored tunnels, toys, and chewable material), which was attributable to TTR up-regulation (109). When these animals were in a heterozygous background for endogeneous TTR (half dose TTR), Aβ levels and deposition were increased in the hippocampus and cortex as compared with their TTR wild-type animals littermates (110). Moreover, TTR neuroprotection in AD was also observed in female APP/PS double transgenic mice bearing a PS A24E mutation; female heterozygous animals for TTR presented increased levels of Aβ42 when compared with TTR wild-type female littermates whereas no differences were found in males of the different genotypes (111). The neuroprotective role of TTR in AD was further corroborated by overexpression of human TTR in the APP23 mouse model (that expresses human APP), leading to decreased Aβ levels and deposition and improved cognition (24).

By contrast, other studies revealed a harmful role of TTR in AD. In the Tg2576 mouse model (transgenic mutant APP), absence of TTR was associated with decreased vascular Aβ burden (112). Moreover, studies in the TgCRND8 mouse strain (strain carrying combined APP mutations), demonstrated that reduction of TTR levels decreased Aβ plaque burden in the hippocampus of 4-month-old animals (113).

All together, the observed effects in the different mice strains, call attention to the interplay of genetic, hormonal and environmental factors known to exist in a complex disease such as AD.

Recently, it was demonstrated that administration of the TTR tetrameric stabilizer iododiflunisal (IDIF) to APP//PS A246E mice, heterozygous for TTR, decreased Aβ deposition and improved cognitive functions (114). Stabilization of TTR through IDIF seems to be a good mechanism to design drugs to AD treatment and sets the basis for further experiments with different TTR tetrameric stabilizers classes.

TTR as an amyloidogenic protein

When the word ‘transthyretin’ is mentioned, the first pathology that emerges is systemic amyloidosis. TTR is the causative agent of a special group of diseases of protein aggregation associated with amyloid fibril deposition. The amyloidoses comprise a spectrum of diseases, either acquired or hereditary, characterized by the extracellular deposition of fibrillar material in specific organs and tissues in the form of amyloid. These fibrils are 7–10 nm wide, rigid, non-branching and are of variable lengths with a typical twisted β-pleated-sheet structure. They present unique tinctorial properties, such as apple-green birefringence under polarized light upon staining with Congo red. The excessive accumulation of these amyloid fibrils gives rise to abnormalities in the function of affected organs. Several apparently nonrelated proteins can be found as main constituents of amyloid fibrils, and this chemical heterogeneity is associated with specific clinical forms of amyloidosis.

Most forms of TTR amyloidosis are hereditary in an autosomal dominant manner; Andrade described the first form of an hereditary amyloidosis in the Northern area of Portugal, near Porto, in family members with age of onset of clinical symptoms in the third to fourth decade of life, nominated familial amyloidotic polyneuropathy (FAP) (115). Early impairment of temperature and pain sensation in the feet, and autonomic dysfunction leading to paresis, malabsorption, emaciation and death were typical clinical features (116, 117). The genetic defect in these Portuguese families has been ascribed to a single point mutation in the TTR gene, originating a mutant TTR (TTR Val30Met) by far, the most frequent TTR mutation associated with FAP in Portugal and elsewhere (118). Over 100 different TTR mutations have been described (119), some of these are associated with FAP clinically, not differing from the original description of the disease; others give rise to variable phenotypes, such as the presence of both neuropathy and cardiomyopathy, presentation of carpal tunnel syndrome, predominant vitreous TTR deposition and leptomeningeal involvement. A few TTR mutations are related to cardiomyopathy without neurological symptoms [familial amyloidotic cardiomyopathy (FAC)]. The most common TTR mutation associated with cardiac amyloidosis is Val122Ile, in the Black population after the age of 60; isolated cardiac amyloidosis is four times more common among Black people than White people in the US and 3.9% of Black people are heterozygous for Val122Ile (120).

TTR can also deposit as amyloid in the heart of elderly people; in the early days this was recognized as a post-mortem finding on 5% of autopsies, a condition termed ‘senile systemic amyloidosis’. With the evolution of imagiology tools, awareness is growing (121).

Normal TTR can also aggregate and form deposits in preeclampsia placental tissue to cause apoptosis. It is present at reduced levels in sera of preeclamptic women, as detected by a proteomic screen (122). The mechanism behind TTR aggregation under this physiological condition is highly unknown.

The precise trigger for TTR aggregation as amyloid is unknown and constitutes a major trend in TTR amyloidosis research. Structural studies correlate the amyloidogenic potential of TTR with weak interactions between subunits. These studies suggest that amyloid fibril formation by some TTR mutants might be triggered by tetramer dissociation to a compact non-native monomer with low conformational stability, which results in partially unfolded monomeric species with a high tendency for ordered aggregation into amyloid fibrils (123). In senile systemic amyloidosis where normal non-mutated TTR deposits in the heart, is has been hypothesized that amyloid formation is related to proteolytic events, or yet unidentified factors related to aging.

Outlook

In the future it is important to dissect in molecular terms the role of TTR in the nervous system, both in the central nervous system (CNS) and peripheral nervous system (PNS) – namely receptors and downstream pathways – to explain the phenotypes so far described in TTR null mice both in normal conditions and in experimental paradigms under stressful conditions. Additionally the role of TTR in other systems, such as the intestinal tract and the placenta should be developed.

Highlights

  • TTR transport of thyroxine and RBP

  • TTR neuroprotection

  • TTR and neurite outgrowth

  • TTR and 14-3-3ζ metabolism

  • TTR in neurodegenerative diseases

  • TTR in behavior

  • Role of TTR in AD and ischemia

  • TTR aggregation.


Corresponding author: Maria João Saraiva, Molecular Neurobiology, Instituto de Biologia Molecular e Celular, IBMC Rua do Campo Alegre, 823, 4150-180 Porto, Portugal; and Instituto de Ciências Biomédicas Abel Salazar, ICBAS, University of Porto, Porto, Portugal, e-mail:

About the authors

Marta Vieira

Marta Vieira (left) has a degree in Biochemistry from the University of Porto and obtained in 2013 her PhD in Biomedical Sciences at the same University participating in several research projects at the Molecular Neurobiology group at the Institute for Cellular and Molecular Biology at Porto University.

Maria João Saraiva

Maria João Saraiva (right) received a BSc in Biology from the University of Porto in 1976, and an MSc in Biochemistry from the University of London, in 1978. Between 1980 and 1984, she studied for a PhD in biochemistry at the University of Porto, and qualified as Professor of Biochemistry in 1991. She worked for different periods as a Visiting Scientist at the College of Physicians and Surgeons at Columbia University, New York. Maria J. Saraiva currently holds a number of positions at the Universidade do Porto, Portugal: she is Vice Director at the Institute for Molecular and Cellular Biology (IBMC); a Professor of Biochemistryat the Biomedical Institute; and Director of the Molecular Neurobiology Group at the IBMC.

References

1. Seibert FB, Nelson JW. Electrophoretic study of the blood protein response in tuberculosis. J Biol Chem 1942; 143: 29–38.10.1016/S0021-9258(18)72655-0Search in Google Scholar

2. Kabat EA, Moore DH, Landow H. An electrophoretic study of the protein components in cerebrospinal fluid and their relationship to the serum proteins. J Clin Invest 1942; 21: 571–7.10.1172/JCI101335Search in Google Scholar

3. Ingbar SH. Pre-albumin: a thyroxine binding protein of human plasma. Endocrinology 1958; 63: 256–9.10.1210/endo-63-2-256Search in Google Scholar

4. Raz A, Goodman DS. The interaction of thyroxine with human plasma prealbumin and with the prealbumin-retinol-binding protein complex. J Biol Chem 1969; 244: 3230–7.10.1016/S0021-9258(18)93118-2Search in Google Scholar

5. Nomenclature committee of the International Union of Biochemistry (NC-IUB). Enzyme nomenclature. Recommendations 1978. Supplement 2: Corrections and additions. Eur J Biochem 1981; 116: 423–35.10.1111/j.1432-1033.1981.tb05353.xSearch in Google Scholar

6. Schreiber G, Richardson SJ. The evolution of gene expression, structure and function of transthyretin. Comp Biochem Physiol B Biochem Mol Biol 1997; 116: 137–60.10.1016/S0305-0491(96)00212-XSearch in Google Scholar

7. Power DM, Elias NP, Richardson SJ, Mendes J, Soares CM, Santos CR. Evolution of the thyroid hormone-binding protein, transthyretin. Gen Comp Endocrinol 2000; 119: 241–55.10.1006/gcen.2000.7520Search in Google Scholar

8. Eneqvist T, Lundberg E, Nilsson L, Abagyan R, Sauer-Eriksson AE. The transthyretin-related protein family. Eur J Biochem 2003; 270: 518–32.10.1046/j.1432-1033.2003.03408.xSearch in Google Scholar

9. Blake CC, Swan ID, Rérat C, Berthou J, Laurent A, Rérat B. An X-ray study of the subunit structure of prealbumin. J Mol Biol 1971; 61: 217–24.10.1016/0022-2836(71)90218-XSearch in Google Scholar

10. Kanda Y, Goodman DS, Canfield RE, Morgan FJ. The amino acid sequence of human plasma prealbumin. J Biol Chem 1974; 249: 6796–805.10.1016/S0021-9258(19)42128-5Search in Google Scholar

11. Blake CC, Geisow MJ, Oatley SJ, Rérat B, Rérat C. Structure of prealbumin: secondary, tertiary and quaternary interactions determined by Fourier refinement at 1.8 Å. J Mol Biol 1978; 121: 339–56.10.1016/0022-2836(78)90368-6Search in Google Scholar

12. Yokoyama T, Mizuguchi M, Nabeshima Y, Kusaka K, Yamada T, Hosoya T, Ohhara T, Kurihara K, Tanaka I, Niimura N. Hydrogen-bond network and pH sensitivity in transthyretin: neutron crystal structure of human transthyretin. J Struct Biol 2012; 177: 283–90.10.1016/j.jsb.2011.12.022Search in Google Scholar

13. Felding P, Fex G. Cellular origin of prealbumin in the rat. Biochim Biophys Acta 1982; 716: 446–9.10.1016/0304-4165(82)90040-XSearch in Google Scholar

14. Aleshire SL, Bradley CA, Richardson LD, Parl FF. Localization of human prealbumin in choroid plexus epithelium. J Histochem Cytochem 1983; 31: 608–12.10.1177/31.5.6341455Search in Google Scholar

15. Smith FR, Goodman DS. The effects of diseases of the liver, thyroid, and kidneys on the transport of vitamin A in human plasma. J Clin Invest 1971; 50: 2426–36.10.1172/JCI106741Search in Google Scholar

16. Stabilini R, Vergani C, Agostoni A, Agostoni RP. Influence of age and sex on prealbumin levels. Clin Chim Acta 1968; 20: 358–9.10.1016/0009-8981(68)90173-3Search in Google Scholar

17. Vahlquist A, Rask L, Peterson PA, Berg T. The concentrations of retinol-binding protein, prealbumin, and transferrin in the sera of newly delivered mothers and children of various ages. Scand J Clin Lab Invest 1975; 35: 569–75.10.3109/00365517509095782Search in Google Scholar

18. Ingenbleek Y, De Visscher M. Hormonal and nutritional status: critical conditions for endemic goiter epidemiology? Metabolism 1979; 28: 9–19.10.1016/0026-0495(79)90162-8Search in Google Scholar

19. Vatassery GT, Quach HT, Smith WE, Benson BA, Eckfeldt JH. A sensitive assay of transthyretin (prealbumin) in human cerebrospinal fluid in nanogram amounts by ELISA. Clin Chim Acta 1991; 197: 19–25.10.1016/0009-8981(91)90344-CSearch in Google Scholar

20. Aldred AR, Brack CM, Schreiber G. The cerebral expression of plasma protein genes in different species. Comp Biochem Physiol B Biochem Mol Biol 1995; 111: 1–15.10.1016/0305-0491(94)00229-NSearch in Google Scholar

21. Schreiber G, Aldred AR, Jaworowski A, Nilsson C, Achen MG, Segal MB. Thyroxine transport from blood to brain via transthyretin synthesis in choroid plexus. Am J Physiol 1990; 258: R338–45.10.1152/ajpregu.1990.258.2.R338Search in Google Scholar

22. Carro E, Trejo JL, Gomez-Isla T, LeRoith D, Torres-Aleman I. Serum insulin-like growth factor I regulates brain amyloid-beta levels. Nat Med 2002; 8: 1390–7.10.1038/nm1202-793Search in Google Scholar

23. Stein TD, Johnson JA. Lack of neurodegeneration in transgenic mice overexpressing mutant amyloid precursor protein is associated with increased levels of transthyretin and the activation of cell survival pathways. J Neurosci 2002; 22: 7380–8.10.1523/JNEUROSCI.22-17-07380.2002Search in Google Scholar

24. Buxbaum JN, Ye Z, Reixach N, Friske L, Levy C, Das P, Golde T, Masliah E, Roberts AR, Bartfai T. Transthyretin protects Alzheimer’s mice from the behavioral and biochemical effects of Abeta toxicity. Proc Natl Acad Sci USA 2008; 105: 2681–6.10.1073/pnas.0712197105Search in Google Scholar

25. Li X, Masliah E, Reixach N, Buxbaum JN. Neuronal production of transthyretin in human and murine Alzheimer’s disease: is it protective? J Neurosci 2011; 31: 12483–90.10.1523/JNEUROSCI.2417-11.2011Search in Google Scholar

26. Sousa JC, Cardoso I, Marques F, Saraiva MJ, Palha JA. Transthyretin and Alzheimer’s disease: where in the brain? Neurobiol Aging 2007; 28: 713–8.10.1016/j.neurobiolaging.2006.03.015Search in Google Scholar

27. Pfeffer BA, Becerra SP, Borst DE, Wong P. Expression of transthyretin and retinol binding protein mRNAs and secretion of transthyretin by cultured monkey retinal pigment epithelium. Mol Vis 2004; 10: 23–30.Search in Google Scholar

28. Kato M, Kato K, Blaner WS, Chertow BS, Goodman DS. Plasma and cellular retinoid-binding proteins and transthyretin (prealbumin) are all localized in the islets of Langerhans in the rat. Proc Natl Acad Sci USA 1985; 82: 2488–92.10.1073/pnas.82.8.2488Search in Google Scholar

29. Jacobsson B, Collins VP, Grimelius L, Pettersson T, Sandstedt B, Carlstrom A. Transthyretin immunoreactivity in human and porcine liver, choroid plexus, and pancreatic islets. J Histochem Cytochem 1989; 37: 31–7.10.1177/37.1.2642294Search in Google Scholar

30. Soprano DR, Herbert J, Soprano KJ, Schon EA, Goodman DS. Demonstration of transthyretin mRNA in the brain and other extrahepatic tissues in the rat. J Biol Chem 1985; 260: 11793–8.10.1016/S0021-9258(17)39100-7Search in Google Scholar

31. Soprano DR, Soprano KJ, Goodman DS. Retinol-binding protein and transthyretin mRNA levels in visceral yolk sac and liver during fetal development in the rat. Proc Natl Acad Sci USA 1986; 83: 7330–4.10.1073/pnas.83.19.7330Search in Google Scholar PubMed PubMed Central

32. Martone RL, Mizuno R, Herbert J. The mammalian pineal gland is a synthetic site for TTR and RBP. J Rheumatol 1993; 20: 175.Search in Google Scholar

33. McKinnon B, Li H, Richard K, Mortimer R. Synthesis of thyroid hormone binding proteins transthyretin and albumin by human trophoblast. J Clin Endocrinol Metab 2005; 90: 6714–20.10.1210/jc.2005-0696Search in Google Scholar

34. Walker AK, Gong Z, Park WM, Zigman JM, Sakata I. Expression of serum retinol binding protein and transthyretin within mouse gastric ghrelin cells. PLoS One 2013; 8: e64882.10.1371/journal.pone.0064882Search in Google Scholar

35. Socolow EL, Woeber KA, Purdy RH, Holloway MT, Ingbar SH. Preparation of I-131-labeled human serum prealbumin and its metabolism in normal and sick patients. J Clin Invest 1965; 44: 1600–9.10.1172/JCI105266Search in Google Scholar

36. Vahlquist A. Metabolism of the vitamin-A-transporting protein complex: turnover of retinol-binding protein, prealbumin and vitamin A in a primate (Macaca Irus). Scand J Clin Lab Invest 1972; 30: 349–60.10.3109/00365517209080270Search in Google Scholar

37. Dickson PW, Howlett GJ, Schreiber G. Metabolism of prealbumin in rats and changes induced by acute inflammation. Eur J Biochem 1982; 129: 289–93.10.1111/j.1432-1033.1982.tb07051.xSearch in Google Scholar

38. Makover A, Moriwaki H, Ramakrishnan R, Saraiva MJ, Blaner WS, Goodman DS. Plasma transthyretin. Tissue sites of degradation and turnover in the rat. J Biol Chem 1988; 263: 8598–603.10.1016/S0021-9258(18)68346-2Search in Google Scholar

39. Sousa MM, Norden AG, Jacobsen C, Willnow TE, Christensen EI, Thakker RV, Verroust PJ, Moestrup SK, Saraiva MJ. Evidence for the role of megalin in renal uptake of transthyretin. J Biol Chem 2000; 275: 38176–81.10.1074/jbc.M002886200Search in Google Scholar

40. Sousa MM, Saraiva MJ. Internalization of transthyretin. Evidence of a novel yet unidentified receptor-associated protein (RAP)-sensitive receptor. J Biol Chem 2001; 276: 14420–5.Search in Google Scholar

41. Mortimer RH, Landers KA, Balakrishnan B, Li H, Mitchell MD, Patel J, Richard K. Secretion and transfer of the thyroid hormone binding protein transthyretin by human placenta. Placenta 2012; 33: 252–6.10.1016/j.placenta.2012.01.006Search in Google Scholar

42. Landers KA, McKinnon BD, Li H, Subramaniam VN, Mortimer RH, Richard K. Carrier-mediated thyroid hormone transport into placenta by placental transthyretin. J Clin Endocrinol Metab 2009; 94: 2610–6.10.1210/jc.2009-0048Search in Google Scholar

43. Bernal J. Thyroid hormones and brain development. Vitam Horm 2005; 71: 95–122.10.1016/S0083-6729(05)71004-9Search in Google Scholar

44. Cheng SY, Leonard JL, Davis PJ. Molecular aspects of thyroid hormone actions. Endocr Rev 2010; 31: 139–70.10.1210/er.2009-0007Search in Google Scholar

45. Loun B, Hage DS. Characterization of thyroxine-albumin binding using high-performance affinity chromatography. I. Interactions at the warfarin and indole sites of albumin. J Chromatogr 1992; 579: 225–35.10.1016/0378-4347(92)80386-5Search in Google Scholar

46. Bartalena L. Recent achievements in studies on thyroid hormone-binding proteins. Endocr Rev 1990; 11: 47–64.10.1210/edrv-11-1-47Search in Google Scholar

47. Hagen GA, Solberg LA, Jr. Brain and cerebrospinal fluid permeability to intravenous thyroid hormones. Endocrinology 1974; 95: 1398–410.10.1210/endo-95-5-1398Search in Google Scholar

48. Chanoine JP, Braverman LE. The role of transthyretin in the transport of thyroid hormone to cerebrospinal fluid and brain. Acta Med Austriaca 1992; 19 Suppl 1:25–8.Search in Google Scholar

49. Blake CC, Geisow MJ, Swan ID, Rerat C, Rerat B. Structure of human plasma prealbumin at 2-5 Å resolution. A preliminary report on the polypeptide chain conformation, quaternary structure and thyroxine binding. J Mol Biol 1974; 88: 1–12.10.1016/0022-2836(74)90291-5Search in Google Scholar

50. Andrea TA, Cavalieri RR, Goldfine ID, Jorgensen EC. Binding of thyroid hormones and analogues to the human plasma protein prealbumin. Biochemistry 1980; 19: 55–63.10.1021/bi00542a009Search in Google Scholar

51. Mendel CM. The free hormone hypothesis: a physiologically based mathematical model. Endocr Rev 1989; 10: 232–74.10.1210/edrv-10-3-232Search in Google Scholar

52. Palha JA, Nissanov J, Fernandes R, Sousa JC, Bertrand L, Dratman MB, Morreale de Escobar G, Gottesman M, Saraiva MJ. Thyroid hormone distribution in the mouse brain: the role of transthyretin. Neuroscience 2002; 113: 837–47.10.1016/S0306-4522(02)00228-2Search in Google Scholar

53. Episkopou V, Maeda S, Nishiguchi S, Shimada K, Gaitanaris GA, Gottesman ME, Robertson EJ. Disruption of the transthyretin gene results in mice with depressed levels of plasma retinol and thyroid hormone. Proc Natl Acad Sci USA 1993; 90: 2375–9.10.1073/pnas.90.6.2375Search in Google Scholar

54. Palha JA, Episkopou V, Maeda S, Shimada K, Gottesman ME, Saraiva MJ. Thyroid hormone metabolism in a transthyretin-null mouse strain. J Biol Chem 1994; 269: 33135–9.10.1016/S0021-9258(20)30107-1Search in Google Scholar

55. Sousa JC, de Escobar GM, Oliveira P, Saraiva MJ, Palha JA. Transthyretin is not necessary for thyroid hormone metabolism in conditions of increased hormone demand. J Endocrinol 2005; 187: 257–66.10.1677/joe.1.06406Search in Google Scholar

56. Mendel CM, Murai JT, Siiter PK, Monroe SE, Inoue M. Conservation of free but not total or non-sex-hormone-binding-globulin-bound testosterone in serum from Nagase analbuminemic rats. Endocrinology 1989; 124(6): 3128–30.10.1210/endo-124-6-3128Search in Google Scholar

57. Refetoff S. Inherited thyroxine-binding globulin abnormalities in man. Endocr Rev 1989; 10: 275–93.10.1210/edrv-10-3-275Search in Google Scholar

58. Bartalena L, Robbins J. Thyroid hormone transport proteins. Clin Lab Med 1993; 13: 583–98.10.1016/S0272-2712(18)30427-XSearch in Google Scholar

59. Patel J, Landers K, Li H, Mortimer RH, Richard K. Delivery of maternal thyroid hormones to the fetus. Trends Endocrinol Metab 2011; 22: 164–70.10.1016/j.tem.2011.02.002Search in Google Scholar PubMed

60. Gudas LJ. Emerging roles for retinoids in regeneration and differentiation in normal and disease states. Biochim Biophys Acta 2012; 1821: 213–21.10.1016/j.bbalip.2011.08.002Search in Google Scholar PubMed PubMed Central

61. Shearer KD, Stoney PN, Morgan PJ, McCaffery PJ. A vitamin for the brain. Trends Neurosci 2012; 35: 733–41.10.1016/j.tins.2012.08.005Search in Google Scholar PubMed

62. Lerner AJ, Gustaw-Rothenberg K, Smyth S, Casadesus G. Retinoids for treatment of Alzheimer’s disease. Biofactors 2012; 38: 84–9.10.1002/biof.196Search in Google Scholar PubMed

63. Corcoran JP, So PL, Maden M. Disruption of the retinoid signalling pathway causes a deposition of amyloid beta in the adult rat brain. Eur J Neurosci 2004; 20: 896–902.10.1111/j.1460-9568.2004.03563.xSearch in Google Scholar PubMed

64. Ono K, Yoshiike Y, Takashima A, Hasegawa K, Naiki H, Yamada M. Vitamin A exhibits potent antiamyloidogenic and fibril-destabilizing effects in vitro. Exp Neurol 2004; 189: 380–92.10.1016/j.expneurol.2004.05.035Search in Google Scholar PubMed

65. Ono K, Yamada M. Vitamin A and Alzheimer’s disease. Geriatr Gerontol Int 2012; 12: 180–8.10.1111/j.1447-0594.2011.00786.xSearch in Google Scholar PubMed

66. Kanai M, Raz A, Goodman DS. Retinol-binding protein: the transport protein for vitamin A in human plasma. J Clin Invest 1968; 47: 2025–44.10.1172/JCI105889Search in Google Scholar PubMed PubMed Central

67. Goodman DS. Vitamin A and retinoids in health and disease. N Engl J Med 1984; 310: 1023–31.10.1056/NEJM198404193101605Search in Google Scholar PubMed

68. Noy N, Slosberg E, Scarlata S. Interactions of retinol with binding proteins: studies with retinol-binding protein and with transthyretin. Biochemistry 1992; 31: 11118–24.10.1021/bi00160a023Search in Google Scholar PubMed

69. Monaco HL, Rizzi M, Coda A. Structure of a complex of two plasma proteins: transthyretin and retinol-binding protein. Science 1995; 268: 1039–41.10.1126/science.7754382Search in Google Scholar PubMed

70. van Bennekum AM, Wei S, Gamble MV, Vogel S, Piantedosi R, Gottesman M, Episkopou V, Blaner WS. Biochemical basis for depressed serum retinol levels in transthyretin-deficient mice. J Biol Chem 2001; 276: 1107–13.10.1074/jbc.M008091200Search in Google Scholar PubMed

71. Wei S, Episkopou V, Piantedosi R, Maeda S, Shimada K, Gottesman ME, Blaner WS. Studies on the metabolism of retinol and retinol-binding protein in transthyretin-deficient mice produced by homologous recombination. J Biol Chem 1995; 270: 866–70.10.1074/jbc.270.2.866Search in Google Scholar PubMed

72. Sundaram M, Sivaprasadarao A, DeSousa MM, Findlay JB. The transfer of retinol from serum retinol-binding protein to cellular retinol-binding protein is mediated by a membrane receptor. J Biol Chem 1998; 273: 3336–42.10.1074/jbc.273.6.3336Search in Google Scholar PubMed

73. Kawaguchi R, Yu J, Honda J, Hu J, Whitelegge J, Ping P, Wiita P, Bok D, Sun H. A membrane receptor for retinol binding protein mediates cellular uptake of vitamin A. Science 2007; 315: 820–5.10.1126/science.1136244Search in Google Scholar PubMed

74. Alapatt P, Guo F, Komanetsky SM, Wang S, Cai J, Sargsyan A, Rodriguez Diaz E, Bacon BT, Aryal P, Graham TE. Liver retinol transporter and receptor for serum retinol-binding protein (RBP4). J Biol Chem 2013; 288: 1250–65.10.1074/jbc.M112.369132Search in Google Scholar PubMed PubMed Central

75. Sousa MM, Berglund L, Saraiva MJ. Transthyretin in high density lipoproteins: association with apolipoprotein A-I. J Lipid Res 2000; 41: 58–65.10.1016/S0022-2275(20)32074-5Search in Google Scholar

76. Liz MA, Faro CJ, Saraiva MJ, Sousa MM. Transthyretin, a new cryptic protease. J Biol Chem 2004; 279: 21431–8.10.1074/jbc.M402212200Search in Google Scholar PubMed

77. Liz MA, Gomes CM, Saraiva MJ, Sousa MM. ApoA-I cleaved by transthyretin has reduced ability to promote cholesterol efflux and increased amyloidogenicity. J Lipid Res 2007; 48: 2385–95.10.1194/jlr.M700158-JLR200Search in Google Scholar PubMed

78. Liz MA, Fleming CE, Nunes AF, Almeida MR, Mar FM, Choe Y, Craik CS, Powers JC, Bogyo M, Sousa MM. Substrate specificity of transthyretin: identification of natural substrates in the nervous system. Biochem J 2009; 419: 467–74.10.1042/BJ20082090Search in Google Scholar PubMed PubMed Central

79. Costa R, Ferreira-da-Silva F, Saraiva MJ, Cardoso I. Transthyretin protects against A-beta peptide toxicity by proteolytic cleavage of the peptide: a mechanism sensitive to the Kunitz protease inhibitor. PLoS One 2008; 3:e2899.10.1371/journal.pone.0002899Search in Google Scholar PubMed PubMed Central

80. Sousa JC, Grandela C, Fernandez-Ruiz J, de Miguel R, de Sousa L, Magalhaes AI, Saraiva MJ, Sousa N, Palha JA. Transthyretin is involved in depression-like behaviour and exploratory activity. J Neurochem 2004; 88: 1052–8.10.1046/j.1471-4159.2003.02309.xSearch in Google Scholar PubMed

81. Prigge ST, Mains RE, Eipper BA, Amzel LM. New insights into copper monooxygenases and peptide amidation: structure, mechanism and function. Cell Mol Life Sci 2000; 57: 1236–59.10.1007/PL00000763Search in Google Scholar

82. Nunes AF, Saraiva MJ, Sousa MM. Transthyretin knockouts are a new mouse model for increased neuropeptide Y. FASEB J 2006; 20: 166–8.10.1096/fj.05-4106fjeSearch in Google Scholar PubMed

83. Sousa JC, Marques F, Dias-Ferreira E, Cerqueira JJ, Sousa N, Palha JA. Transthyretin influences spatial reference memory. Neurobiol Learn Mem 2007; 88: 381–5.10.1016/j.nlm.2007.07.006Search in Google Scholar PubMed

84. Brouillette J, Quirion R. Transthyretin: a key gene involved in the maintenance of memory capacities during aging. Neurobiol Aging 2008; 29: 1721–32.10.1016/j.neurobiolaging.2007.04.007Search in Google Scholar PubMed

85. Fleming CE, Saraiva MJ, Sousa MM. Transthyretin enhances nerve regeneration. J Neurochem 2007; 103: 831–9.10.1111/j.1471-4159.2007.04828.xSearch in Google Scholar PubMed

86. Fleming CE, Mar FM, Franquinho F, Saraiva MJ, Sousa MM. Transthyretin internalization by sensory neurons is megalin mediated and necessary for its neuritogenic activity. J Neurosci 2009; 29: 3220–32.10.1523/JNEUROSCI.6012-08.2009Search in Google Scholar PubMed PubMed Central

87. Dekki N, Refai E, Holmberg R, Kohler M, Jornvall H, Berggren PO, Juntti-Berggren L. Transthyretin binds to glucose-regulated proteins and is subjected to endocytosis by the pancreatic beta-cell. Cell Mol Life Sci 2012; 69: 1733–43.10.1007/s00018-011-0899-8Search in Google Scholar PubMed

88. Refai E, Dekki N, Yang SN, Imreh G, Cabrera O, Yu L, Yang G, Norgren S, Rossner SM, Inverardi L, Ricordi C, Olivecrona G, Andersson M, Jornvall H, Berggren PO, Juntti-Berggren L. Transthyretin constitutes a functional component in pancreatic beta-cell stimulus-secretion coupling. Proc Natl Acad Sci USA 2005; 102: 17020–5.10.1073/pnas.0503219102Search in Google Scholar PubMed PubMed Central

89. Vieira M, Saraiva MJ. Transthyretin regulates hippocampal 14-3-3zeta protein levels. FEBS Lett 2013; 587: 1482–8.10.1016/j.febslet.2013.03.011Search in Google Scholar PubMed

90. Nunes AF, Montero M, Franquinho F, Santos SD, Malva J, Zimmer J, Sousa MM. Transthyretin knockout mice display decreased susceptibility to AMPA-induced neurodegeneration. Neurochem Int 2009; 55: 454–7.10.1016/j.neuint.2009.07.001Search in Google Scholar PubMed

91. Zerr I, Bodemer M, Gefeller O, Otto M, Poser S, Wiltfang J, Windl O, Kretzschmar HA, Weber T. Detection of 14-3-3 protein in the cerebrospinal fluid supports the diagnosis of Creutzfeldt-Jakob disease. Ann Neurol 1998; 43: 32–40.10.1002/ana.410430109Search in Google Scholar PubMed

92. Gloeckner SF, Meyne F, Wagner F, Heinemann U, Krasnianski A, Meissner B, Zerr I. Quantitative analysis of transthyretin, tau and amyloid-beta in patients with dementia. J Alzheimers Dis 2008; 14: 17–25.10.3233/JAD-2008-14102Search in Google Scholar PubMed

93. Zhang HL, Zhang XM, Mao XJ, Deng H, Li HF, Press R, Fredrikson S, Zhu J. Altered cerebrospinal fluid index of prealbumin, fibrinogen, and haptoglobin in patients with Guillain-Barre syndrome and chronic inflammatory demyelinating polyneuropathy. Acta Neurol Scand 2012; 125: 129–35.10.1111/j.1600-0404.2011.01511.xSearch in Google Scholar PubMed

94. Ruetschi U, Zetterberg H, Podust VN, Gottfries J, Li S, Hviid Simonsen A, McGuire J, Karlsson M, Rymo L, Davies H, Minthon L, Blennow K. Identification of CSF biomarkers for frontotemporal dementia using SELDI-TOF. Exp Neurol 2005; 196: 273–81.10.1016/j.expneurol.2005.08.002Search in Google Scholar PubMed

95. Brettschneider J, Lehmensiek V, Mogel H, Pfeifle M, Dorst J, Hendrich C, Ludolph AC, Tumani H. Proteome analysis reveals candidate markers of disease progression in amyotrophic lateral sclerosis (ALS). Neurosci Lett 2010; 468: 23–7.10.1016/j.neulet.2009.10.053Search in Google Scholar PubMed

96. Arguelles S, Venero JL, Garcia-Rodriguez S, Tomas-Camardiel M, Ayala A, Cano J, Machado A. Use of haptoglobin and transthyretin as potential biomarkers for the preclinical diagnosis of Parkinson’s disease. Neurochem Int 2010; 57: 227–34.10.1016/j.neuint.2010.05.014Search in Google Scholar PubMed

97. Chen R, Vendrell I, Chen CP, Cash D, O’Toole KG, Williams SA, Jones C, Preston JE, Wheeler JX. Proteomic analysis of rat plasma following transient focal cerebral ischemia. Biomark Med 2011; 5: 837–46.10.2217/bmm.11.89Search in Google Scholar PubMed

98. Suzuyama K, Shiraishi T, Oishi T, Ueda S, Okamoto H, Furuta M, Mineta T, Tabuchi K. Combined proteomic approach with SELDI-TOF-MS and peptide mass fingerprinting identified the rapid increase of monomeric transthyretin in rat cerebrospinal fluid after transient focal cerebral ischemia. Brain Res Mol Brain Res 2004; 129: 44–53.10.1016/j.molbrainres.2004.06.021Search in Google Scholar PubMed

99. Liverman CS, Cui L, Yong C, Choudhuri R, Klein RM, Welch KM, Berman NE. Response of the brain to oligemia: gene expression, c-Fos, and Nrf2 localization. Brain Res Mol Brain Res 2004; 126: 57–66.10.1016/j.molbrainres.2004.02.028Search in Google Scholar PubMed

100. Santos SD, Lambertsen KL, Clausen BH, Akinc A, Alvarez R, Finsen B, Saraiva MJ. CSF transthyretin neuroprotection in a mouse model of brain ischemia. J Neurochem 2010; 115: 1434–44.10.1111/j.1471-4159.2010.07047.xSearch in Google Scholar PubMed

101. Hornstrup LS, Frikke-Schmidt R, Nordestgaard BG, Tybjaerg-Hansen A. Genetic stabilization of transthyretin, cerebrovascular disease, and life expectancy. Arterioscler Thromb Vasc Biol 2013; 33: 1441–7.10.1161/ATVBAHA.113.301273Search in Google Scholar PubMed

102. Goedert M, Spillantini MG. A century of Alzheimer’s disease. Science 2006; 314: 777–81.10.1126/science.1132814Search in Google Scholar PubMed

103. Elovaara I, Maury CP, Palo J. Serum amyloid A protein, albumin and prealbumin in Alzheimer’s disease and in demented patients with Down’s syndrome. Acta Neurol Scand 1986; 74: 245–50.10.1111/j.1600-0404.1986.tb07863.xSearch in Google Scholar PubMed

104. Castano EM, Roher AE, Esh CL, Kokjohn TA, Beach T. Comparative proteomics of cerebrospinal fluid in neuropathologically-confirmed Alzheimer’s disease and non-demented elderly subjects. Neurol Res 2006; 28: 155–63.10.1179/016164106X98035Search in Google Scholar PubMed

105. Han SH, Jung ES, Sohn JH, Hong HJ, Hong HS, Kim JW, Na DL, Kim M, Kim H, Ha HJ, Kim YH, Huh N, Jung MW, Mook-Jung I. Human serum transthyretin levels correlate inversely with Alzheimer’s disease. J Alzheimers Dis 2011; 25: 77–84.10.3233/JAD-2011-102145Search in Google Scholar PubMed

106. Ribeiro CA, Santana I, Oliveira C, Baldeiras I, Moreira J, Saraiva MJ, Cardoso I. Transthyretin decrease in plasma of MCI and AD patients: investigation of mechanisms for disease modulation. Curr Alzheimer Res 2012; 9: 881–9.10.2174/156720512803251057Search in Google Scholar PubMed

107. Schwarzman AL, Gregori L, Vitek MP, Lyubski S, Strittmatter WJ, Enghilde JJ, Bhasin R, Silverman J, Weisgraber KH, Coyle PK, Zagorski MG, Tafous J, Eisenberg M, Saunders AM, Roses AD, Goldgaber D. Transthyretin sequesters amyloid beta protein and prevents amyloid formation. Proc Natl Acad Sci USA 1994; 91: 8368–72.10.1073/pnas.91.18.8368Search in Google Scholar PubMed PubMed Central

108. Liu L, Murphy RM. Kinetics of inhibition of beta-amyloid aggregation by transthyretin. Biochemistry 2006; 45: 15702–9.10.1021/bi0618520Search in Google Scholar PubMed

109. Lazarov O, Robinson J, Tang YP, Hairston IS, Korade-Mirnics Z, Lee VM, Hersh LB, Sapolsky RM, Mirnics K, Sisodia SS. Environmental enrichment reduces Aβ levels and amyloid deposition in transgenic mice. Cell 2005; 120: 701–13.10.1016/j.cell.2005.01.015Search in Google Scholar PubMed

110. Choi SH, Leight SN, Lee VM, Li T, Wong PC, Johnson JA, Saraiva MJ, Sisodia SS. Accelerated Abeta deposition in APPswe/PS1ΔE9 mice with hemizygous deletions of TTR (transthyretin). J Neurosci 2007; 27: 7006–10.10.1523/JNEUROSCI.1919-07.2007Search in Google Scholar PubMed PubMed Central

111. Oliveira SM, Ribeiro CA, Cardoso I, Saraiva MJ. Gender-dependent transthyretin modulation of brain amyloid-beta levels: evidence from a mouse model of Alzheimer’s disease. J Alzheimers Dis 2011; 27: 429–39.10.3233/JAD-2011-110488Search in Google Scholar PubMed

112. Wati H, Kawarabayashi T, Matsubara E, Kasai A, Hirasawa T, Kubota T, Harigaya Y, Shoji M, Maeda S. Transthyretin accelerates vascular Abeta deposition in a mouse model of Alzheimer’s disease. Brain Pathol 2009; 19: 48–57.10.1111/j.1750-3639.2008.00166.xSearch in Google Scholar PubMed PubMed Central

113. Doggui S, Brouillette J, Chabot JG, Farso M, Quirion R. Possible involvement of transthyretin in hippocampal β-amyloid burden and learning behaviors in a mouse model of Alzheimer’s disease (TgCRND8). Neurodegener Dis 2010; 7: 88–95.10.1159/000285513Search in Google Scholar PubMed

114. Ribeiro CA, Oliveira SM, Guido LF, Magalhaes A, Valencia G, Arsequell G, Saraiva MJ, Cardoso I. Transthyretin stabilization by iododiflunisal promotes amyloid-β peptide clearance, decreases its deposition, and ameliorates cognitive deficits in an Alzheimer’s disease mouse model. J Alzheimers Dis 2014; 39(2): 357–70.10.3233/JAD-131355Search in Google Scholar PubMed

115. Andrade C. A peculiar form of peripheral neuropathy; familiar atypical generalized amyloidosis with special involvement of the peripheral nerves. Brain 1952; 75: 408–27.10.1093/brain/75.3.408Search in Google Scholar PubMed

116. Dyck PJ, Lambert EH. Dissociated sensation in amylidosis. Compound action potential, quantitative histologic and teased-fiber, and electron microscopic studies of sural nerve biopsies. Arch Neurol 1969; 20: 490–507.10.1001/archneur.1969.00480110054005Search in Google Scholar PubMed

117. Sousa MM, Saraiva MJ. Neurodegeneration in familial amyloid polyneuropathy: from pathology to molecular signaling. Prog Neurobiol 2003; 71: 385–400.10.1016/j.pneurobio.2003.11.002Search in Google Scholar PubMed

118. Coelho T, Maurer MS, Suhr OB. THAOS – The transthyretin amyloidosis outcomes survey: initial report on clinical manifestations in patients with hereditary and wild-type transthyretin amyloidosis. Curr Med Res Opin 2013; 29: 63–76.10.1185/03007995.2012.754348Search in Google Scholar PubMed

119. Sekijima Y, Yoshida K, Tokuda T, Ikeda S. Familial Transthyretin Amyloidosis. 2001 [updated January 26, 2012]; Available from: http://www.ncbi.nlm.nih.gov/books/NBK1194/.Search in Google Scholar

120. Jacobson DR, Pastore RD, Yaghoubian R, Kane I, Gallo G, Buck FS, Buxbaum JN. Variant-sequence transthyretin (isoleucine 122) in late-onset cardiac amyloidosis in black Americans. N Engl J Med 1997; 336: 466–73.10.1056/NEJM199702133360703Search in Google Scholar PubMed

121. Westermark P, Sletten K, Johansson B, Cornwell GG, 3rd. Fibril in senile systemic amyloidosis is derived from normal transthyretin. Proc Natl Acad Sci USA 1990; 87: 2843–5.10.1073/pnas.87.7.2843Search in Google Scholar PubMed PubMed Central

122. Kalkunte SS, Neubeck S, Norris WE, Cheng SB, Kostadinov S, Vu Hoang D, Ahmed A, von Eggeling F, Shaikh Z, Padbury J, Berg G, Olofsson A, Markert UR, Sharma S. Transthyretin is dysregulated in preeclampsia, and its native form prevents the onset of disease in a preclinical mouse model. Am J Pathol 2013; 183: 1425–36.10.1016/j.ajpath.2013.07.022Search in Google Scholar PubMed PubMed Central

123. Quintas A, Vaz DC, Cardoso I, Saraiva MJ, Brito RM. Tetramer dissociation and monomer partial unfolding precedes protofibril formation in amyloidogenic transthyretin variants. J Biol Chem 2001; 276: 27207–13.10.1074/jbc.M101024200Search in Google Scholar PubMed

Received: 2013-12-9
Accepted: 2014-1-15
Published Online: 2014-02-12
Published in Print: 2014-03-01

©2014 by Walter de Gruyter Berlin/Boston

This article is distributed under the terms of the Creative Commons Attribution Non-Commercial License, which permits unrestricted non-commercial use, distribution, and reproduction in any medium, provided the original work is properly cited.

Downloaded on 28.3.2024 from https://www.degruyter.com/document/doi/10.1515/bmc-2013-0038/html
Scroll to top button