This Event is licensed under the Creative Commons BY-SA license. This license allows reusers to distribute, remix, adapt, and build upon the material in any medium or format, so long as attribution is given to the creator. The license allows for commercial use. If you remix, adapt, or build upon the material, you must license the modified material under identical terms.

Event: 960

Key Event Title

A descriptive phrase which defines a discrete biological change that can be measured. More help

Increased, Uptake of thyroxine into tissue

Short name
The KE short name should be a reasonable abbreviation of the KE title and is used in labelling this object throughout the AOP-Wiki. More help
Increased, Uptake of thyroxine into tissue
Explore in a Third Party Tool

Biological Context

Structured terms, selected from a drop-down menu, are used to identify the level of biological organization for each KE. More help
Level of Biological Organization
Cellular

Cell term

The location/biological environment in which the event takes place.The biological context describes the location/biological environment in which the event takes place.  For molecular/cellular events this would include the cellular context (if known), organ context, and species/life stage/sex for which the event is relevant. For tissue/organ events cellular context is not applicable.  For individual/population events, the organ context is not applicable.  Further information on Event Components and Biological Context may be viewed on the attached pdf. More help

Organ term

The location/biological environment in which the event takes place.The biological context describes the location/biological environment in which the event takes place.  For molecular/cellular events this would include the cellular context (if known), organ context, and species/life stage/sex for which the event is relevant. For tissue/organ events cellular context is not applicable.  For individual/population events, the organ context is not applicable.  Further information on Event Components and Biological Context may be viewed on the attached pdf. More help

Key Event Components

The KE, as defined by a set structured ontology terms consisting of a biological process, object, and action with each term originating from one of 14 biological ontologies (Ives, et al., 2017; https://aopwiki.org/info_pages/2/info_linked_pages/7#List). Biological process describes dynamics of the underlying biological system (e.g., receptor signalling).Biological process describes dynamics of the underlying biological system (e.g., receptor signaling).  The biological object is the subject of the perturbation (e.g., a specific biological receptor that is activated or inhibited). Action represents the direction of perturbation of this system (generally increased or decreased; e.g., ‘decreased’ in the case of a receptor that is inhibited to indicate a decrease in the signaling by that receptor).  Note that when editing Event Components, clicking an existing Event Component from the Suggestions menu will autopopulate these fields, along with their source ID and description.  To clear any fields before submitting the event component, use the 'Clear process,' 'Clear object,' or 'Clear action' buttons.  If a desired term does not exist, a new term request may be made via Term Requests.  Event components may not be edited; to edit an event component, remove the existing event component and create a new one using the terms that you wish to add.  Further information on Event Components and Biological Context may be viewed on the attached pdf. More help
Process Object Action
transport thyroxine increased

Key Event Overview

AOPs Including This Key Event

All of the AOPs that are linked to this KE will automatically be listed in this subsection. This table can be particularly useful for derivation of AOP networks including the KE.Clicking on the name of the AOP will bring you to the individual page for that AOP. More help
AOP Name Role of event in AOP Point of Contact Author Status OECD Status
Transthyretin interference KeyEvent Kristie Sullivan (send email) Under Development: Contributions and Comments Welcome Under Development

Taxonomic Applicability

Latin or common names of a species or broader taxonomic grouping (e.g., class, order, family) that help to define the biological applicability domain of the KE.In many cases, individual species identified in these structured fields will be those for which the strongest evidence used in constructing the AOP was available in relation to this KE. More help

Life Stages

An indication of the the relevant life stage(s) for this KE. More help

Sex Applicability

An indication of the the relevant sex for this KE. More help

Key Event Description

A description of the biological state being observed or measured, the biological compartment in which it is measured, and its general role in the biology should be provided. More help

T4 (and T3) is actively transported across the cell membrane into target tissues through the action of specific carrier-mediated uptake (simple diffusion probably plays a minor role), where it is T3 that binds to and triggers the nuclear receptors in the target cells (Yen 2001, Zoeller et al 2007). The T3 supply is met via secretion from the thyroid (20%) and through conversion of T4 into T3 (80%) through the action of outer-ring deiodinase enzymes D1 and D3 (Chopra 1996 from Zoeller et al 2007). THs are cleared from serum by the liver following sulfation (via sulfotransferase enzymes) or glucuronidation (via UDP-glucuronosyl transferase enzymes) and ultimately, eliminated in the bile (Hood and Klaasen 2000).

Two major groups of transporters have been identified: organic anionic transport proteins (OATPs) and amino acid transporters (L- and T-type). Several of these transporters have displayed greater affinity and selectivity for T4 and T3 and specific compounds (such as polychlorinated biphenyls and polybrominated diethyl ethers) have been found to bind to the transport proteins either in serum or in various cellular compartments (Zoeller et al 2007)

OATPs transport both iodothyronines as well as sulfated conjugates and the gene family (SLCO) coding for this family of homologous proteins is clustered on human 12p12 (Hagenbuch and Meier 2004). OATP1A2 is expressed in brain, liver and kidney while OATP1B1, -1B2, and -1B3 are expressed in the liver and display high affinity for both T4 and T3 (Friesma et al 2005). OATP1C1 shows binding preference for T4 over T3 and is almost exclusively expressed in brain capillaries, where it is thought to play a role for transport of T4 across the blood brain barrier (Tohyama et al 2004).

There is also evidence that L-type or T-type amino acid transport proteins also play a role in cellular uptake of thyroid hormones. The former transport large neutral branched-chain and aromatic amino acids while the latter are specific to the aromatic amino acids Phe, Tyr and Trp (Visser 2010). The T-type amino acid transport protein TAT1 has been cloned from both rats and humans, is encoded by SLC16A10, and is a member of the monocarboxylate transporter (MCT) family (MCT10)(Kim et al 2002). Both MCT and MCT8 share a high degree of homology and are both highly effective iodothyronine transporters. MCT8 is highly selective for T4 and T, responsible for transporting T3 into neuronal cells and interferes with brain development if absent (Friesma et al 2003, 2006 and 2008). MCT8 is expressed in multiple tissues, including liver, kidney, heart, brain, placenta, thyroid, skeletal muscle and adrenal gland, while MCT10 is also expressed in various tissues, with high expression in muscle, intestine, kidney and pancreas and the former is known to only transport iodothyronine molecules while the latter can also carry aromatic amino acids (Nishimura et al 2008). In terms of transport efficacy, MCT10 appears to be superior to MCT8 for moving T3; however, the reverse is true for T4.

Uptake into hepatocytes is probably mediated through multiple low-affinity/high-capacity and high-affinity/low-capacity processes that can be inhibited by certain molecules, such as fatty acids, bilirubin and indoxyl sulfate (Henneman et al 2001). Km values for these processes are in the micro- to nanomolar range, while the serum concentrations of free T4 and T3 are in the picomolar range.

Accumulation in tissue as well as transport across the placental and blood-brain barrier is dependent on expression level of the various TH membrane transporters, such as those in the MCT and OATP families. MCT8, MCT10, L-type amino acid transporters LAT1 and LAT2, OATP1A2 and OATP4A1 are responsible for transplacental transfer of TH (Loubiere et al 2010), while MCT8 and OATP1C1 are important to transfer across the blood-brain barrier (Roberts et al 2008).

OATP1B1 and OATP1B3 show preference for sulfated THs and are expressed only in the liver

OATP1C1 shows high preference for T4 and almost exclusively expressed in brain capillaries and choroid plexus (Hagenbuch 2007)

LAT1 and LAT2 facilitate bidirectional transport of both T4 and T3 (and aliphatic and aromatic amino acids) across the plasma membrane

MCT8 only transports THs and expressed in choroid plexus, MCT10 also transports aromatic amino acids

Schroder van der Elst et al 1997 injected female rats with synthetic flavonoid [125I]-EMD 49209 and [131I]-T4 (or rats pretreated with EMD 21388), noting rapid clearance of [125I] from serum and rapid uptake of [131I] into tissues. These results show that the flavonoid itself does not cross the blood-brain barrier, despite the fact that they demonstrate displacement of T4 from TTR and temporarily increase the pool of available free T4. 

Schroder van der Elst et al 1998 injected pregnant rats at GD 20 with [125I]-EMD 49209 and observed distribution in maternal tissues, intestinal contents and fetal tissues. No flavonoid was detected in the brain but it was found in all fetal tissues examined, including brain.

TRANSPORT ACROSS THE BLOOD-BRAIN BARRIER

Active transport is required for uptake of T3 and T4 across cell membranes (Heuer 2007). Accumulation in tissue as well as transport across the placental and blood-brain barrier (BBB) is dependent on expression level of the various TH membrane transporters, such as those in the MCT and OATP families. MCT8, MCT10, L-type amino acid transporters LAT1 and LAT2, OATP1A2 and OATP4A1 are responsible for transplacental transfer of TH (Loubiere et al 2010), while MCT8 and OATP1C1 are important to transfer across the blood-brain barrier (Roberts et al 2008).  Visser et al 2008 proposed that OATP14 is mainly responsible for moving T4 into the brain/CSF, where it is converted into T3 locally and transported into neurons via MCT8.

Schreiber et al 1995; Schreiber 2002

Roberts et al 2008 examined the expression of MCT8 and OATP14 in male Sprague Dawley rats, male CD-1 mice and human brain tissue via qPCR and immunofluorescent staining as well staining and confocal microscopy in isolated cerebral microvessels and choroid plexus (CP) epithelium.  They observed that the main transporter at the BBB is OATP14 whereas MCT8 mediates TH uptake into neuronal cells.  MCT8 mRNA and proteins were expressed in cerebral microvessels in all species; however, OATP14 mRNA and protein was only enriched in mouse and rat microvessels.  In all species, MCT8 is concentrated on the epithelial cell apical surface and OATP14 primarily on the basal-lateral surface of the CP epithelial cells.  These data suggested MCT8 plays a role in TH transport across the BBB and this is supported by the pattern of localization of the two transporters.

Kim et al 2015 screened protein transporters in rat serum for the potential to guide nanoparticles across the BBB (via receptor-mediated transcytosis, or RMT) using an in vitro transcytosis assay using rat and human brain microvascular endothelial cells.  Eleven (11) proteins were identified as showing potential to penetrate the endothelial cell layer via RMT, including Ttr.  Ttr was then incorporated into a quantum dot nanoparticle, administered to male Sprague-Dawley rats via IV and found to cross the BBB in rats via transcytosis, confirmed by in vivo imaging, TEM, ICP-MS and confocal microscopy.

TRANSPORT ACROSS PLACENTA

There is a direct role for maternal TH in the development of the fetal CNS starting with the 1st trimester and this maternal TH must be provided to the fetus via transplacental delivery (Chan et al 2002; de Escobar et al 2004).  The placenta responds to TH with both the villous and extravillous trophoblasts (EVTs) expressing specific nuclear receptor isoforms for T3.  The primary barrier of cells for maternal-fetal exchange are the syncytiotrophoblasts of placental villi, which are in direct contact with maternal blood, and the cytotrophoblasts, which form an additional inner layer of cells (Benirschke et al 2000).  Free T4 is the believed to be the primary TH transported across the placenta and fetal free T4 levels reach ~ 40-50% of maternal concentrations by the early 2nd trimester and peak in the early 3rd trimester, where they remain at levels higher than the corresponding maternal concentration (Calvo et al 2002; Hume et al 2004).

Accumulation in tissue as well as transport across the placental and blood-brain barrier is dependent on expression level of the various TH membrane transporters, such as those in the MCT and OATP families. MCT8, MCT10, L-type amino acid transporters LAT1 and LAT2, OATP1A2 and OATP4A1 are responsible for transplacental transfer of TH (Loubiere et al 2010), while MCT8 and OATP1C1 are important to transfer across the blood-brain barrier (Roberts et al 2008).  The MCT8, OATP4A1 and LAT1 are localized at the apical membrane of the syncytiotrophoblasts while MCT10 is localized in the cytotrophoblasts during the 1st trimester (Ritchie and Taylor 2001; Sato et al 2003; Chan et al 2006; Loubiere et al 2010)

Displacement of T4 from transport proteins during the developmental stage could have consequences for both fetal development and later in adulthood (Morse et al 1996).  Transfer of maternal TH across the placenta is essential to neurodevelopment and even temporary disruption during the perinatal period can have long-term adverse health effects (Zoeller and Rovet 2004). Animal studies have confirmed that perinatal exposure to PBDEs can adversely affect neurodevelopment of CNS; however, the mechanisms remain elusive and the evidence in humans that hydroxylated metabolites of PBDEs is equivocal (Costa et al 2014).

TTR mediates transport through the placenta and the hydroxylated metabolites of PCBs and PBDEs have been found to more potent than the natural ligand T4 and thus, competitive binders. The presence of these compounds in maternal and infant blood have been associated with changes in TH, developmental endpoints and fertility in humans (Chevrier et al 2010, Harley et al 2010, Koopman-Esseboom et al 1994).  Several studies have demonstrated links between decreased T4 levels and neurodevelopmental and neurobehavioral adverse outcomes in mice exposed specifically to BDE-99 (Branchi et al 2002; Viberg et al 2002).

Darnerud et al 1996 treated pregnant C57BL and NMRI mice on GD 13 with a single gavage at two doses of [14C]-labelled 3,3’,4,4’-tetraCB (PCB 77) and experiment was terminated after 4 days, measuring radioactivity and TH in maternal and fetal liver and plasma. Competitive binding assay was also done with [125I]-T4 complex from samples of fetal and maternal plasma. Dose-dependent uptake of [14C] were noted in both maternal and fetal plasma and liver, with fetal plasma radioactivity levels being 4- to 9-fold higher than maternal levels and corresponded to a single metabolite (4-OH-tetraCB).  Gel electrophoresis confirmed the [14C] was bound to fetal serum TTR and the fetal sera samples at the top dose (10 mg/kg) showed 50% TTR binding relative to controls, along with significant decrease in free and total serum T4.

Morse et al 1996 (also see Morse et al 1993) treated Wistar rats with Aroclor 1254 at 2 doses via daily oral exposure from GD 10 to 16, with blood and tissue (brain, liver) collected from dams and fetuses on GD 20 with pups reared until PND Day 21 (with blood and tissue collected at PND 4 and 21).  The biological samples were analyzed for TH, type II deiodinase activity and levels of PCBs and metabolites.  Maternal exposure to Aroclor 1254 significantly reduced both free and total T4 in the serum of fetus and neonate (Day 4) in a dose-dependent manner (but less pronounced at PND 21 and absent at Day 90). At GD 20, levels of T4 in fetal forebrain and cerebellum, and at PND 21 female weanlings at the high dose (25 mg/kg) still had significantly decreased forebrain T4.  In the fetus, deiodination (T4 to T3) was significantly increased in forebrain and in the female weanling, only at the low dose was deiodination significantly decreased. Similarly, glucuronidation was significantly decreased in the fetus and significantly increased in the female weanling. Accumulation of mainly one metabolite (2,3,3’,4’,5-pentachloro-4-biphenylol, or 4-OH-pentaCB; possibly from PCB 118 or PCB 126) was noted in fetal serum and forebrain as well as neonatal and weanling plasma and the concentrations of the metabolite in plasma relative to the more persistent parent congeners (PCB 153) were increased all the way to 90 days (and the plasma levels of the offspring exceeded that of dam all the way to 90 days).  These data show that maternal exposure to PCBs can result in accumulation of hydroxylated metabolites in fetal plasma that reduces T4 and, as a result, reduces brain levels of T4 with a compensatory increase in brain deiodination to maintain brain T3 concentration.

Pedraza et al 1996 treated pregnant Wistar rats first with methimazole (to block hormone synthesis) and then continuous infusion of EMD 21388 and T4 from GD 11 to 21, noting decreased total T4, increased free T4 and decreased T3 in maternal serum, increase of T3 in placenta and led to measurable amounts of parent compound in fetal serum along with decreased total T4 and increased T3.

Schroder van der Elst et al 1998 injected pregnant rats at GD 20 with [125I]-EMD 49209 and observed distribution in maternal tissues, intestinal contents and fetal tissues. No flavonoid was detected in the maternal brain but it was found in all fetal tissues examined, including brain.  It should be noted though that TTR is the principal carrier in the fetal rat and the EMD flavonoids were designed as T4 analogs and only bind to TTR (and not to albumin or TBG).  Shortly after birth, TTR production decreases to nearly zero and thus, interference with TTR-T4 during certain developmental windows might impact availability of thyroid hormone in certain tissues at critical time periods.

Sinjari and Darnerud (1998) injected C57BL mice on GD 16 with 5 doses of [14C]-labelled metabolites of PCB 77 (4-OH-tetraCB, two different 4-OH-pentaCB metabolites of PCB105), sacrificed 24 hours later, plasma and tissues collected from dam and fetus and analyzed for [14C], total T4 and liver microsomes.  Partial dose dependency was found for both maternal and fetal decreased total T4 for 4-OH-tetraCB and one of the pentaCBs.  Also, placental transfer to fetal plasma was dose dependent and, at lower doses (less than 5 mg/kg), fetal serum levels of 4-OH-tetraCB were 2-fold higher than maternal serum levels.  The authors conclude that doses in excess of 5 mg/kg saturate ligand binding, as effects measured at concentrations higher than this are not dose-related (however, there was extensive biliary excretion of 4-OH-tetraCB at the highest doses (20 and 50 µmol/kg).  These results suggest that hydroxylated metabolites of PCBs are transferred to fetus upon maternal exposure, but did not induce a CYP1A1 or CYP1A2 response in the dam and competitive binding with T4 may not be the only mechanism behind noted adverse fetal effects from T4 modulation.

Meerts et al 2002 treated pregnant rats with 5 mg/kg of 4-OH-CB107 (radiolabeled and non-labelled) on GD 10 to 16, noting accumulation in the fetal compartment. The complex between TTR and [14C]-4-OH-CB107 was detected in serum in both dam and fetus.  Total and free serum T4 were reduced in fetus at GD 17 and 20. T4 concentration in fetal forebrain homogenate was reduced at GD20 and deiodination of T4 to T3 was increased at GD 17.  No changes were noted in maternal or fetal hepatic UDP-UGT activity, type 1 deiodination, or EROD activity.  These data show that TTR-mediated transport of xenobiotics, like the metabolites of PCB 107, can result in transfer from mother to fetus and can result in reduced fetal T4 (although it should be noted that fetal T3 remained unaffected). Furthermore, there was significant increase of fetal TSH at GD20, indicating stimulation of the HPT axis.

Inoue et al 2004 reported that PFOS can cross the placental barrier in humans

McKinnon et al 2005

Morse et al 2005 injected pregnant Wistar rats on GD 13 with single dose of [14C]-labelled 3,3’,4,4’-tetraCB (PCB 77) and tracked metabolism for 7 days.  The main metabolite was 4-OH-CB77 was found in maternal liver and plasma, placental tissue and fetal plasma, with 4-OH-CB77 accumulating 100-fold in fetus over the observation period with levels in fetal plasma being 14-fold higher than maternal plasma on GD 20. Similarly, while maternal serum T4 was initially significantly reduced and recovered by GD 20, the fetal plasma T4 was found to be significantly reduced relative to maternal T4 on GD 20. These data show that exposure to PCBs during pregnancy can result on transfer of metabolites to fetus (and these are competitive with T4 at the TTR binding site).

Riu et al 2008 fed pregnant Wistar rats with [14C]-decabromodiphenyl ether (DBDE) over 96 hr of late gestation (GD 16 to 19) and tissues analyzed via HPLC.  More than 19% of the administered dose was recovered, with 2/3 of this eliminated via feces, and results in accordance with past findings in Fisher and Sprague Dawley rats.  Small amounts were found to cross both the blood-brain and placental barriers and hydroxylated octaBDE was found in all tissues and fetus.

Dalliare et al 2009a looked at the relationship between TH status TBG and exposure to PCB-153, pentachlorophenol, hexachlorobenzene and hydroxylated PCBs in pregnant Inuit women and their infants. PCB-153 was the most predominant congener found to be elevated most in pregnant women, followed by infant and cord plasma levels.  OH-PCB results were a sum of 11 major congeners and found to be higher in pregnant women than cord blood, but highly intercorrelated.  Overall, the results suggest that the compounds measured in serum were not significant predictors of TH or TSH concentrations in this population.  The strongest results were found for PCP, which was negatively associated with free T4 in neonate cord blood, suggesting PCP reduces the transfer of T4 across the placenta. This confirmed previous findings of Sandau et al 2002, but has conflicted with other study populations and published reports; however, has biological plausibility as PCP has been reported to have a binding affinity twice that of the natural ligand for TTR (van den Berg 1990).

Loubiere et al 2010 described the ontogeny of TH transporters MCT8, MCT10, LAT1, LAT2, OATP1A2 and OATP4A1 in over 100 placenta samples collected across gestation via RNA extraction and qRT-PCR.  These mRNA data showed increasing expression of MCT8, MCT10, OATP1A2 and LAT1 throughout gestation, while OATP4A1 and CD98 (associated with LAT activity) mRNA fell to a nadir in the late 1st and early 2nd trimester. Immunohistochemistry data localized MCT10 and OATP1A2 for the first time to EVTs as well as syncytiotrophoblasts.

How It Is Measured or Detected

A description of the type(s) of measurements that can be employed to evaluate the KE and the relative level of scientific confidence in those measurements.These can range from citation of specific validated test guidelines, citation of specific methods published in the peer reviewed literature, or outlines of a general protocol or approach (e.g., a protein may be measured by ELISA). Do not provide detailed protocols. More help

Total T4 is most often measured using human serum based diagnostic kits, but free T4 (and T3) is only directly measured through equilibrium dialysis and ultrafiltration (Midgley 2001). Large volumes of serum must be used due to the very low concentrations of free T4 normally found (0.1% of total T4), which requires pooling of samples taken from fetus or pup. Some researchers have tried to “micronize” this process through combining RIA to measure total TH and dialysis to estimate the free fraction (Zoeller et al 2007).  Extracted materials can also be quantified by HPLC. The reference range for free T4 is 9.8 to 18.8 pM/L (Dirinck et al 2016).

T3 is found in similar plasma concentrations to T4 (i.e. 5-10 pM) with < 0.4% being in the unbound state. Measuring free serum T3 is labor intensive and requires equipment not available in many clinical reference laboratories and thus ultrafiltration is often used (Abdalla and Bianco 2014). Immunoassays and MS/MS are also used.

Measuring displacement of T4 from serum transport proteins is done mainly via one of three in vitro methods: radioligand binding assay, plasmon resonance-based biosensor, or fluorescence displacement.

Radioligand binding assays, using [125I]-T4 as a label, were developed to demonstrate affinity for xenobiotics to human or rat TTR and TBG (Brouwer and van den Berg 1986, Lans et al 1994).  The most commonly used method was first published by Somack et al 1982 and adapted by Hamers et al 2006, Lans et al 1993 and Ucan-Marin et al 2010. Similar assays have been developed using [125I]-T3 as a label for affinity to chicken and bullfrog TTR (Yamauchi et al 2003).  Radioligand methods suffer from having to use heavily regulated isotopes and lower throughput to provide free T4 measurements (due to the extra wash/separation procedure needed). The most well-known protocol uses TTR purified from human serum (which may not be as stable as recombinant) and performed in a pure aqueous solution, which may not be as stable for lipophilic compounds (Chauhan et al 2000 is an example using PCBs).

Purkey et al 2001 published a binding assay using polyclonal TTR antibodies covalently bound to sepharose resin which is then mixed with plasma pre-treated with compound of interest, washed and analyzed via HPLC.

Marchesini et al 2006 reported on the development of two surface plasmon resonance(SPR)-based biosensor assays using recombinant TTR and TBG, validated with known thyroid disruptors and structurally related compounds including halogenated phenols, polychlorinated biphenyls, bisphenols and a hydroxylated PCB metabolite (4-OH-CB 14).  TH is covalently bound to a gold-layered chip and a mixture of the compound of interest and transport protein are injected in a flow cell passing over the bound TH.  The authors found that these biosensor methods were more sensitive (IC50 of 8.6 ± 0.7 nM for rTTR), easier to perform and more rapid that radioligand binding assays and immunoprecipitation-HPLC.

Marchesini et al 2008 applied their biosensor-based screen to 62 chemicals of public health concern and found that hydroxylated metabolites of PCBs (particularly para-hydroxylated ones) and PBDEs (BDEs 47, 49 and 99) displayed the most potent binding to TBG and TTR, confirming many other previous studies.  The authors conclude their optimized assays are suitable for high-throughput screening for potential thyroid disruption.

Cao et al 2010, Cao et al 2011 and Ren and Guo 2012 developed the FLU-TTR, based on a protein-binding fluorescent probe (ANSA, or 8-anilo-1-naphthalenesulfonic acid ammonium salt) that becomes highly fluorescent after binding to T4. When the compound of interest is introduced and displaces the ANSA-thyroxine probe, this fluorescence is reduced.  This allows generation of binding constant (K) data as opposed to past efforts that generated IC50 values.  Cao et al 2011 developed a fluorescent microtiter method for pTTR and TBG tested with bisphenol A.

Montano et al 2012 developed a competitive T4-TTR fluorescence displacement assay in a 96-well format, modified from the original method (Nilsson and Petersen 1975) and using a new selective method to extract hydroxylated metabolites while reducing fatty acid interference (modified from Hovander et al 2000).

Aqai et al 2012 described a rapid and isotope-free (13C6-T4) screening of thyroid transport protein ligands, using a competitive binding assay for rTTR using fast ultrahigh performance LC-electrospray ionization triple-quadrupole MS. The method involves the use of immunomagnetic beads followed by screening with flow cytometry and UPLC-MS. The high-throughput screening mode is capable of detecting T4 in water at the part-per-trillion level and in the part-per-billion level in urine.

Relevant Phase II enzymes that are responsible for TH metabolism include UGT1A1, UGT1A6 and SULT2A1 while relevant cellular import/export transport proteins include MCT8, OATP1A4 and MRP2. All contribute towards systemic clearance of TH and conjugates from serum whether increasing biliary excretion or moving TH into tissues and across the placenta and BBB.  Enzyme induction can only be measured via in vitro cell-based assays and since these enzymes are all controlled by specific nuclear receptors, assays targeting these receptors might act as surrogate measurement (Murk et al 2013). Several methods measuring expression of UGT or SULT mRNA have been published; however, there have been limited efforts to develop higher-throughput methods.  The EPA ToxCast Phase I efforts used quantitative nuclease protection assays (qNPA) to screen several hundred chemicals for UGT1A1 and SULT2A1 (Rotroff et al 2010, Sinz et al 2006).

Domain of Applicability

A description of the scientific basis for the indicated domains of applicability and the WoE calls (if provided).  More help

References

List of the literature that was cited for this KE description. More help

Abdalla, S.M. and A.C. Bianco. (2014) Defending plasma T3 is a biological priority.  Clin. Endocrinol. (Oxf)  81(5): 633-641.

Alshehri, B., D’Souza, D. G., Lee, J. Y., Petratos, S., & Richardson, S. J. (2015). The Diversity of Mechanisms Influenced by Transthyretin in Neurobiology: Development, Disease and Endocrine Disruption. Journal of Neuroendocrinology, 27(5), 303–323. http://doi.org/10.1111/jne.12271

Andrea, T.A., R.R. Cavalieri, I.D. Goldfine and E.C. Jorgensen (1980) Binding of thyroid hormones and analogues to the human plasma protein prealbumin. Biochemistry  19(1): 55-63.

Aqai, P., C. Fryganas, M. Mizuguchi, W. Haasnoot and M.W. Nielen. (2012) Triple bioaffinity mass spectrometry concept for thyroid transporter ligands.  Anal. Chem.  84(15): 6488-6493.

Athanasiadou, M., S.N. Cuadra, G. Marsh, A> Bergman, and K. Jakobsson. (2008) Polybrominated diphenyl ethers (PBDEs) and bioaccumulative hydroxylated PBDE metabolites in young humans from Managua, Nicaragua.  Environ. Health Perspect. 116(3): 400-408.

Barter, R.A. and C.D. Klaassen. (1994) Reduction of thyroid hormone levels and alteration of thyroid function by four representative UDP-glucuronosyltransferase inducers in rats.  Toxicol. Appl. Pharmacol.  128(1): 9-17.

Blake, C.C., J.M. Burridge and S.J. Oatley. (1978) X-ray analysis of thyroid hormone binding to prealbumin. Biochem Soc. Trans. 6(6): 1114-1118.

Bloom, M.S., J.E. Vena, J.R. Olson and P.J. Kostyniak.  (2009)  Assessment of polychlorinated biphenyl congeners, thyroid stimulating hormone, and free thyroxine among New York state anglers.  Int. J. Hyg. Environ. Health  212(6): 599-611.

Branchi, I., E. Alleva and L.G. Costa.  (2002)  Effects of perinatal exposure to a polybrominated diphenyl ether (PBDE 99) on mouse neurobehavioural development.  Neurotoxicology  23(3): 375-384.

Brouwer, a, & van den Berg, K. J. (1986). Binding of a metabolite of 3,4,3’,4'-tetrachlorobiphenyl to transthyretin reduces serum vitamin A transport by inhibiting the formation of the protein complex carrying both retinol and thyroxin. Toxicology and Applied Pharmacology, 85(3), 301–312.

Calvo, R.M., E. Jauniaux, B. Gulbis, M. Asuncion, C. Gervy, B. Contempre and G. Morreale de Escobar.  (2002)  Fetal tissues are exposed to biologically relevant free thyroxine concentrations during early phases of development.  J. Clin. Endocrinol. Metab.  87(4); 1768-1777.

Cao, J., L.H. Guo, B. Wan and Y. Wei. (2011) In vitro fluorescence displacement investigation of thyroxine transport disruption by bisphenol A.  J. Environ Sci, (China)  23(2): 315-321.

Cao, J., Y. Lin, L.H. Guo, A.Q. Zhang, Y. Wei and Y. Yang. (2010) Structure-based investigation on the binding interaction of hydroxylated polybrominated diphenyl ethers with thyroxine transport proteins.  Toxicology  277(1-3): 20-28.

Chan, S.Y., J.A. Franklyn, H.N. Pemberton, J.N. Bulmer, T.J. Visser, C.J. McCabe and M.D. Kilby.  (2006)  Monocarboxylate transporter 8 expression in the human placenta: the effects of severe intrauterine growth restriction.  J. Endocrinol.  189(3): 465-471.

Chan, S., S. Kachilele, C.J. McCabe, L.A. Tannahill, K. Boelaert, N.J. Gittoes, T.J. Visser, J.A. Franklyn and M.D. Kilby.  (2002)  Early expression of thyroid hormone deiodinases and receptors in human fetal cerebral cortex.  Brain Res. Dev. Brain Res.  138(2): 109-116.

Chang, S.C., J.R. Thibodeaux, M.L. Eastvold, D.J. Ehresman, J.A. Bjork, J.W. Froehlich, C. Lau, R.J. Singh, K.B. Wallace and J.L. Butenhoff. (2008) Thyroid hormone status and pituitary function in adult rats given oral doses of perfluorooctanesulfonate (PFOS).  Toxicology  243(3): 330-339.

Chanoine, J.-P., Alex, S., Fang, S. L., Stone, S., Leonard, J. L., Kohrle, J., & Braverman, L. E. (1992). Role of transthyretin in the transport of thyroxine from the blood to the choroid plexus, the cerebrospinal fluid and the brain. Endocrinology, 130(2), 933–938.

Chauhan, K. R., Kodavanti, P. R. S., & McKinney, J. D. (2000). Assessing the Role of ortho-Substitution on Polychlorinated Biphenyl Binding to Transthyretin, a Thyroxine Transport Protein. Toxicology and Applied Pharmacology, 162(1), 10–21. http://doi.org/10.1006/taap.1999.8826

Cheek, A.O., K. Kow, J. Chen and J.A. McLachlan. (1999) Potential mechanisms of thyroid disruption in humans: interaction of organochlorine compounds with thyroid receptor, transthyretin, and thyroid-binding globulin.  Environ. Health Perspect.  107(4): 273-278.

Chevrier, J., K.G. Harley, A. Bradman, M. Gharbi, A. Sjodin and B. Eskenazi.  (2010)  Polybrominated diphenyl ether (PBDE) flame retardants and thyroid hormone during pregnancy.  Environ. Health Perspect.  118(10) : 1444-1449.

Chopra, I.J., P. Taing and L. Mikus. (1996) Direct determination of free triiodothyronine (T3) in undiluted serum by equilibrium dialysis/radioimmunoassay (RIA).  Thyroid  6(4): 255-259.

Costa, L.G., R. de Laat, S. Tagliaferri and C. Pellacani.  (2014)  A mechanistic view of polybrominated diphenyl ether (PBDE) developmental neurotoxicity.  230(2): 282-294.

Dallaire, R., G. Muckle, E. Dewailly, S.W. Jacobson, J.L. Jacobson, T.M. Sandanger, C.D. Sandau and P. Ayotte. (2009a)  Thyroid hormone levels of pregnant inuit women and their infants exposed to environmental contaminants.  Environ. Health Perspect.  117(6): 1014-1020.

Dallaire, R., E. Dewailly, D. Pereg, S. Dery and P. Ayotte.  (2009b)  Thyroid function and plasma concentrations of polyhalogenated compounds in Inuit adults.  Environ. Health Perspect.  117(9): 1380-1386.

Darnerud, P.O., D. Morse, E. Klasson-Wehler and A Brouwer.  (1996)  Binding of a 3,3', 4,4'-tetrachlorobiphenyl (CB-77) metabolite to fetal transthyretin and effects on fetal thyroid hormone levels in mice.  Toxicology  106(1-3): 105-114.

De Escobar, G.M., M.J. Obregon and F.E. del Rey.  (2004)  Maternal thyroid hormones early in pregnancy and fetal brain development.  Best Pract. Res. Clin. Endocrinol. Metab.  18(2): 225-248.

Dirinck, E., A.C. Dirtu, G. Malarvanna, A. Covaci, P.G. Jorens and L.F. Van Gall. (2016) A Preliminary Link between Hydroxylated Metabolites of Polychlorinated Biphenyls and Free Thyroxin in Humans.  Int. J. Environ. Res. Public Health  13(4): 421.

Eguchi, A., K. Nomiyama, N. Minh Tue, P.T. Trang, P. Hung Viet, S. Takahashi and S. Tanabe.  (2015)  Residue profiles of organohalogen compounds in human serum from e-waste recycling sites in North Vietnam: Association with thyroid hormone levels.  Environ. Res.  137: 440-449.

Emerson, C.H., J.H. Cohen III, R.A Yung, S. Alex and S.L. Fang. (1990) Gender-related differences of serum thyroxine-binding proteins in the rat. Acta Endocrinol. (Copenh)  123(1): 72-78.

Erratico, C.A., A. Steitz and S.M. Bandiera. (2013) Biotransformation of 2,2',4,4'-tetrabromodiphenyl ether (BDE-47) by human liver microsomes: identification of cytochrome P450 2B6 as the major enzyme involved.  Chem. Res. Toxicol.  26(5): 721-731.

Erratico, C.A., S.C. Moffatt and S.M. Bandiera. (2011) Comparative oxidative metabolism of BDE-47 and BDE-99 by rat hepatic microsomes.  Toxicol. Sci.  123(1): 37-47.

Eskenazi, B., J. Chevrier, S.A. Rauch, K. Kogul, K.G. Harley, C. Johnson, C. Trujillo, A. Sjodin and A. Bradman.  (2013)  In utero and childhood polybrominated diphenyl ether (PBDE) exposures and neurodevelopment in the CHAMACOS study.  121(2) : 257-262.

Feo, M.L., M.S. Gross, B.P. McGarrigle, E. Eljarrat, D. Barcelo, D.S. Aga and J.R. Olson. (2013) Biotransformation of BDE-47 to potentially toxic metabolites is predominantly mediated by human CYP2B6.  Environ. Health Persepct.  121(4): 440-446.

Ferguson, R.N., H. Edelhoch, H.A. Saroff, J. Robbins and H.J. Cahnmann (1975) Negative cooperativity in the binding of thyroxine to human serum prealbumin. Preparation of tritium-labeled 8-anilino-1-naphthalenesulfonic acid.  Biochemistry  14(2): 282-289.

Friesema EC, Jansen J, Jachtenberg JW, Visser WE, Kester MH, Visser TJ 2008 Effective cellular uptake and efflux of thyroid hormone by human monocarboxylate transporter 10. Molecular endocrinology (Baltimore, Md 22:1357-1369

Friesema EC, Kuiper GG, Jansen J, Visser TJ, Kester MH 2006 Thyroid hormone transport by the human monocarboxylate transporter 8 and its rate-limiting role in intracellular metabolism. Molecular endocrinology (Baltimore, Md 20:2761-2772

Friesma, E.C., J. Jansen and T.J. Visser. (2005) Thyroid hormone transporters.  Biochem. Soc. Trans.  33(part 1): 228-232.

Friesema EC, Ganguly S, Abdalla A, Manning Fox JE, Halestrap AP, Visser TJ 2003 Identification of monocarboxylate transporter 8 as a specific thyroid hormone transporter. J Biol Chem 278:40128-40135

Grimm, F. a., Lehmler, H. J., He, X., Robertson, L. W., & Duffel, M. W. (2013). Sulfated metabolites of polychlorinated biphenyls are high-affinity ligands for the thyroid hormone transport protein transthyretin. Environmental Health Perspectives, 121(6), 657–662.

Gutshall, D.M., G.D. Pilcher and A.E. Langley. (1989) Mechanism of the serum thyroid hormone lowering effect of perfluoro-n-decanoic acid (PFDA) in rats. J. Toxicol. Environ. Health   28(1): 53-65.

Hagenbuch, B. (2007)  Cellular entry of thyroid hormones by organic anion transporting polypeptides.  Best Pract. Res. Clin. Endocrinol. Metab.  21(2): 209-221.

Hagenbuch B, Meier PJ 2004 Organic anion transporting polypeptides of the OATP/ SLC21 family: phylogenetic classification as OATP/ SLCO superfamily, new nomenclature and molecular/functional properties. Pflugers Arch 447:653-665

Hagmar, L., L. Rylander, E. Dyremark, E. Klasson-Wehler and E.M. Erfurth. (2001a).  Plasma concentrations of persistent organochlorines in relation to thyrotropin and thyroid hormone levels in women.  Int. Arch. Occup. Environ. Health  74(3): 184-188.

Hagmar, L., J. Bjork, A. Sjodin, A. Bergman and E.M. Erfurth. (2001b) Plasma levels of persistent organohalogens and hormone levels in adult male humans.  Arch. Environ. Health  56(2): 138-143.

Hallgren, S., T. Sinjari, H. Hakansson and P.O. Darnerud. (2001) Effects of polybrominated diphenyl ethers (PBDEs) and polychlorinated biphenyls (PCBs) on thyroid hormone and vitamin A levels in rats and mice.  75(4): 200-208.

Hallgren, S. and P.O. Darnerud. (2002) Polybrominated diphenyl ethers (PBDEs), polychlorinated biphenyls (PCBs) and chlorinated paraffins (CPs) in rats-testing interactions and mechanisms for thyroid hormone effects.  Toxicology  177(203): 227-243.

Hamers, T., J.H. Kamstra, E. Sonneveld, A.J. Murk, M.H. Kester, P.L. Andersson, J. Legler and A. Brouwer. (2006) In vitro profiling of the endocrine-disrupting potency of brominated flame retardants.  Toxicol. Sci.  92(1): 157-173.

Hamers, T., Kamstra, E. Sonneveld, A.J. Murk, T.J. Visser, M.J. Van Velzen, A. Brouwer and A. Bergman. (2008) Biotransformation of brominated flame retardants into potentially endocrine-disrupting metabolites, with special attention to 2,2',4,4'-tetrabromodiphenyl ether (BDE-47).  Mol. Nutr. Food Res.  52(2): 284-298.

Harley, K.G., A.R. Marks, J. Chevrier, A. Bradman, A. Sjodin and B. Eskenazi.  (2010)  PBDE concentrations in women's serum and fecundability.  Environ. Health Perspect.  118(5): 699-704.

Henneman, G., R. Docter, E.C. Friesma, M. de Jong, E.P. Krenning and T.J. Visser.  (2001)  Plasma membrane transport of thyroid hormones and its role in thyroid hormone metabolism and bioavailability.  Endocr. Rev.  22(4): 451-476.

Heuer, H.  (2007)  The importance of thyroid hormone transporters for brain development and function.  Best Pract. Res. Clin. Endocrinol. Metab.  21(2):  265-276.

Hood, A. and C.D. Klaassen. (2000a) Differential effects of microsomal enzyme inducers on in vitro thyroxine (T(4)) and triiodothyronine (T(3)) glucuronidation.  Toxicol. Sci.  55(1): 78-84.

Hood, A. and C.D. Klaassen.  (2000b)  Effects of microsomal enzyme inducers on outer-ring deiodinase activity toward thyroid hormones in various rat tissues.  Toxicol. Appl. Pharmacol.  163(3): 240-248.

Hovander, L., M. Athanasiadou, L. Asplund, S. Jensen and E.K. Wehler. (2000). Extraction and cleanup methods for analysis of phenolic and neutral organohalogens in plasma.  24(8): 696-703.

Hume, R., J. Simpson, C. Delahunty, H. van Toor, S.Y. Wu, F.L. Williams, T.J. Visser et al.  (2004) Human fetal and cord serum thyroid hormones: developmental trends and interrelationships.  J. Clin. Endocrinol. Metab.  89(8): 4097-4103.

Inoue, K., F. Okada, R. Ito, S. Kato, S. Sasaki, S. Nakajima, A. Uno, Y. Saijo, F. Sata, Y. Yoshimura, R. Kishi and H. Nakazawa. (2004) Perfluorooctane sulfonate (PFOS) and related perfluorinated compounds in human maternal and cord blood samples: assessment of PFOS exposure in a susceptible population during pregnancy.  Environ. Health Perspect.  112(11): 1204-1207.

Kato, Y., K. Haraguchi, M. Onishi, S. Ikushiro, T. Endo, C. Ohta, N. Koga, S Yamada and M. Degawa. (2014) 3,3',4,4'-Tetrachlorobiphenyl-mediated decrease of serum thyroxine level in C57BL/6 and DBA/2 mice occurs mainly through enhanced accumulation of thyroxine in the liver.  Biol. Pharm. Bull.  37(3) 504-509.

Kato, Y., M. Onishi, K. Haraguchi, S. Ikushiro, C. Ohta, N. Koga, T. Endo, S. Yamada and M. Degawa. (2013) A possible mechanism for 2,3',4,4',5'-pentachlorobiphenyl-mediated decrease in serum thyroxine level in mice.  Biol. Pharm. Bull.  36(10): 1594-1601.

Kato, Y., S. Tamaki, K. Haraguchi, S. Ikushiro, M. Sekimoto, C. Ohta, T. Endo, N. Koga, S. Yamada and M. Degawa. (2012) Comparative study on 2,2',4,5,5'-pentachlorobiphenyl-mediated decrease in serum thyroxine level between C57BL/6 and its transthyretin-deficient mice.  Toxicol. Appl. Pharmacol.  263(3): 323-329.

Kato, Y., M. Onishi, K. Haraguchi, S. Ikushiro, C. Ohta, N. Koga, T. Endo, S. Yamada and M. Degawa. (2011) A possible mechanism for 2,2',4,4',5,5'-hexachlorobiphenyl-mediated decrease in serum thyroxine level in mice.  Toxicol. Appl. Pharmacol.  254(1): 48-55.

Kato, Y., K. Haraguchi, M. Kubota, Y. Seto, S. Ikushiro, T. Sakaki, N. Koga, S. Yamada and M. Degawa. (2009) 4-Hydroxy-2,2',3,4',5,5',6-heptachlorobiphenyl-mediated decrease in serum thyroxine level in mice occurs through increase in accumulation of thyroxine in the liver.  Drug Metab. Dispos.  37(10): 2095-2102.

Kato, Y., S. Ikushiro, R. Takiguchi, K. Haraguchi, N. Koga, S. Uchida, T. Sakaki, S. Yamada, J. Kanno and M. Degawa. (2007) A novel mechanism for polychlorinated biphenyl-induced decrease in serum thyroxine level in rats.  Drug Metab. Dispos. 35(10) : 1949-1955.

Kato, Y., S. Ikushiro, K. Haraguchi, T. Yamazaki, Y. Ito, H. Suzuki, R. Kimura, S. Yamada, T. Inoue and M. Degawa. (2004) A possible mechanism for decrease in serum thyroxine level by polychlorinated biphenyls in Wistar and Gunn rats.  Toxicol. Sci.  81(2): 309-315.

Kato, Y., K. Haraguchi, T. Yamazuki, Y. Ito, S. Miyajima, K. Nemoto, N. Koga, R. Kimura and M. Degawa. (2003) Effects of polychlorinated biphenyls, kanechlor-500, on serum thyroid hormone levels in rats and mice.  Toxicol. Sci.  72(2): 235-241.

Kim, S.Y., E.S. Choi, H.J. Lee, C. Moon and E. Kim.  (2015)  Transthyretin as a new transporter of nanoparticles for receptor-mediated transcytosis in rat brain microvessels.  Colloids Surf B Biointerfaces  136: 989-996.

Kim do K, Kanai Y, Matsuo H, Kim JY, Chairoungdua A, Kobayashi Y, Enomoto A, Cha SH, Goya T, Endou H 2002 The human T-type amino acid transporter-1: characterization, gene organization, and chromosomal location. Genomics 79:95-103

Kohrle, J., S.L. Fang, Y. Yang, K. Irmscher, R.D. Hesch, S. Pino, S. Alex, and L.E. Braverman. (1989). Rapid effects of the flavonoid EMD 21388 on serum thyroid hormone binding and thyrotropin regulation in the rat. Endocrinoloy 125: 532-537

Koopman-Essenboom, C., D.C. Morse, N. Weisglas-Kuperus, I.J. Lutkeschipholt, C.G. Van der Paauw, L.G. Tuinstra, A. Brouwer and P.J. Sauer.  (1994)  Effects of dioxins and polychlorinated biphenyls on thyroid hormone status of pregnant women and their infants.  Pediatr. Res.  36(4): 468-473.

Lans, M. C., Klasson-Wehler, E., Willemsen, M., Meussen, E., Safe, S., & Brouwer, A. (1993). STRUCTURE-DEPENDENT, COMPETITIVE INTERACTION OF HYDROXY-POLYCHLOROBIPHENYLS, -DIBENZO-p-DIOXINS AND -DIBENZOFURANS WITH HUMAN TRANSTHYRETIN. Chemico-Biological Interactions, 88, 7–21.

Lans, M. C., Spiertz, C., Brouwer, a, & Koeman, J. H. (1994). Different competition of thyroxine binding to transthyretin and thyroxine-binding globulin by hydroxy-PCBs, PCDDs and PCDFs. European Journal of Pharmacology, 270(2-3), 129–136. http://doi.org/10.1016/0926-6917(94)90054-X

Larsson, M., Pettersson, T., & Carlström, a. (1985). Thyroid hormone binding in serum of 15 vertebrate species: isolation of thyroxine-binding globulin and prealbumin analogs. General and Comparative Endocrinology, 58(3), 360–375.

Loubiere, L.S., E. Vasilopoulou, J.N. Bulmer, P.M. Taylor, B. Stieger, F. Verrey, C.J. McCabe, J.A. Franklyn, M.D. Kilby and S.Y. Chan. (2010)  Expression of thyroid hormone transporters in the human placenta and changes associated with intrauterine growth restriction.  Placenta  31(4): 295-304.

Lueprasitsakul, W., Alex, S., Fang, S. L., Pino, S., Irmscher, K., Köhrle, J., & Braverman, L. E. (1990). Flavonoid administration immediately displaces thyroxine (T4) from serum transthyretin, increases serum free T4, and decreases serum thyrotropin in the rat. Endocrinology 126 (6)

Lupton, S.J., P. McGarrigle, J.R. Olson, T.D. Wood and D.S. Aga. (2010) Analysis of hydroxylated polybrominated diphenyl ether metabolites by liquid chromatography/atmospheric pressure chemical ionization tandem mass spectrometry.  Rapid Commun. Mass. Spectrom.  24(15): 2227-2235.

Lupton, S.J., B.P. McGarrigle, J.R. Olson, T.D. Wood and D.S. Aga. (2009)  Analysis of hydroxylated polybrominated diphenyl ether metabolites by liquid chromatography/atmospheric pressure chemical ionization tandem mass spectrometry.  22(11): 1802-1809.

Malmberg, T., M. Athanasiadou, G. Marsh, I. Brandt and A. Bergman.  (2005) Identification of hydroxylated polybrominated diphenyl ether metabolites in blood plasma from polybrominated diphenyl ether exposed rats.  39(14): 5342-5348.

Marchesini, G.R., E. Meulenberg, W. Haasnoot, M. Mizuguchi and H. Irth.  (2006) Biosensor recognition of thyroid-disrupting chemicals using transport proteins.  Anal. Chem.  78(4): 1107-1114.

Marchesini, G.R., A. Meimaridou, W. Haasnoot, E. Meulenberg, F. Albertus, M. Mizuguchi, M. Takeuchi, H. Irth and A.J. Murk. (2008) iosensor discovery of thyroxine transport disrupting chemicals.  Toxicol. Appl. Pharmacol.  232(1): 150-160.

Martin, L.A., D.T. Wilson, K.R> Reuhl, M.A. Gallo and C.D. Klaassen. (2012) Polychlorinated biphenyl congeners that increase the glucuronidation and biliary excretion of thyroxine are distinct from the congeners that enhance the serum disappearance of thyroxine.  Drug Metab. Dispos.  40(3): 588-595.

Martin, L. and C.D. Klaassen. (2010) Differential effects of polychlorinated biphenyl congeners on serum thyroid hormone levels in rats.  Toxicol. Sci.  117(1): 36-44.

Meerts, I.A., Y. Assink, P.H. Cenjin, J.H. Van Den Berg, B.M. Weijers, A. Bergman, J.H. Koeman and A. Brouwer. (2002) Placental transfer of a hydroxylated polychlorinated biphenyl and effects on fetal and maternal thyroid hormone homeostasis in the rat.  Toxicol. Sci. 68(2): 361-371.

Meerts, I.A., J.J. van Zanden, E.A. Lujiks, I. van Leeuwen-Bol, G. Marsh, E. Jakobsson, A. Bergman and A. Brouwer. (2000) Potent competitive interactions of some brominated flame retardants and related compounds with human transthyretin in vitro.  Toxicol. Sci.  56(1): 95-104.

Mendel, C. M. (1989). Modeling thyroxine transport to liver : rejection of the “enhanced dissociation” hypothesis as applied to thyroxine. Am J Physiol, 257(Endocrinol Metab 20), E764–E771.

Mendel, C. M., Cavalieri, R. R., & Kohrle, J. (1992). Thyroxine (T4) transport and distribution in rats treated with EMD 21388, a synthetic flavonoid that displaces T4 from transthyretin. Endocrinology, 130(3), 1525–1532.

Midgley, J. E. (2001) Direct and indirect free thyroxine assay methods: theory and practice.  Clin. Chem.  47(8): 1353-1363.

Miksys, S. and R.F. Tyndale. (2004) The unique regulation of brain cytochrome P450 2 (CYP2) family enzymes by drugs and genetics.  Drug Metab. Rev.  36(2): 313-333.

Montano, M., E. Coccco, C. Guignard, G. Marsh, L. Hoffmann, A. Bergman, A.C. Gutleb and A.J. Murk. (2012) New approaches to assess the transthyretin binding capacity of bioactivated thyroid hormone disruptors.  Toxicol. Sci.  130(1): 94-105.

Morse, D.C., E.K. Wehler, W. Wesseling, J.H. Koeman and A. Brouwer.  (1996)  Alterations in rat brain thyroid hormone status following pre- and postnatal exposure to polychlorinated biphenyls (Aroclor 1254).  Toxicol. Appl. Pharmacol.  136(2): 269-279.

Morse, D.C., D. Groen, M. Veerman, C.J. van Amerongen, H.B. Koeter, A.E. Smits van Proojie, T.J. Visser, J.H. Koeman and A. Brouwer.  (1993)  Interference of polychlorinated biphenyls in hepatic and brain thyroid hormone metabolism in fetal and neonatal rats.  Toxicol. Appl. Pharmacol.  122(1) :27-33.

Munro, S.L., C.F. Lim, J.G. Hall, J.W. Barlow, D.J. Craik, D.J. Topliss and J.R. Stockigt (1989) Drug competition for thyroxine binding to transthyretin (prealbumin): comparison with effects on thyroxine-binding globulin. J. Clin. Endocrinol. Metab.  68(6): 1141-1147,

Nishimura M, Naito S 2008 Tissue-specific mRNA expression profiles of human solute carrier transporter superfamilies. Drug Metab Pharmacokinet 23:22-44

Pedraza, P., Calvo, R., Obregón, M. J., Asuncion, M., Escobar Del Rey, F., & Morreale De Escobar, G. (1996). Displacement of T4 from transthyretin by the synthetic flavonoid EMD 21388 results in increased production of T3 from T4 in rat dams and fetuses. Endocrinology, 137(11), 4902–4914. http://doi.org/10.1210/en.137.11.4902

Purkey, H.E., M.I. Dorrell and J.W. Kelly. (2001) Evaluating the binding selectivity of transthyretin amyloid fibril inhibitors in blood plasma.  Proc. Natl. Acad. Sci. USA  98(10): 5566-5571.

Refetoff, S., N.I. Robin and V.S. Fang. (1970) Parameters of thyroid function in serum of 16 selected vertebrate species: a study of PBI, serum T4, free T4, and the pattern of T4 and T3 binding to serum proteins.  Endocrinology  86(4): 793-805.

Refetoff, S. (2015) Thyroid Hormone Serum Transport Proteins. In: De Groot LJ, Chrousos G, Dungan K, Feingold KR, Grossman A, Hershman JM, Koch C, Korbonits M, McLachlan R, New M, Purnell J, Rebar R, Singer F, Vinik A, editors. Endotext [Internet]. South Dartmouth (MA): MDText.com, Inc.; 2000.

Ren, X.M., L.H. Guo, Y. Gao, B.T. Zhang and B. Wan. (2013) Hydroxylated polybrominated diphenyl ethers exhibit different activities on thyroid hormone receptors depending on their degree of bromination.  Toxicol. Appl. Pharamacol.  268(3): 256-263.

Ren, X. M., & Guo, L. H. (2012). Assessment of the binding of hydroxylated polybrominated diphenyl ethers to thyroid hormone transport proteins using a site-specific fluorescence probe. Environmental Science and Technology, 46(8), 4633–4640. http://doi.org/10.1021/es2046074

Rerat, C. and H.G. Schwick (1967) [Crystallographic data of blood plasma prealbumin]. [Article in French] Acta Crystallogr.  22(3): 441-442.

Richardson, S. J. (2007). Cell and molecular biology of transthyretin and thyroid hormones. International Review of Cytology, 258(January), 137–93. http://doi.org/10.1016/S0074-7696(07)58003-4

Richardson, S. J., Wijayagunaratne, R. C., D’Souza, D. G., Darras, V. M., & Van Herck, S. L. J. (2015). Transport of thyroid hormones via the choroid plexus into the brain: the roles of transthyretin and thyroid hormone transmembrane transporters. Frontiers in Neuroscience, 9(March), 1–8.

Rickenbacher, U., McKinney, J. D., Oatley, S. J., & Blake, C. C. (1986). Structurally specific binding of halogenated biphenyls to thyroxine transport protein. Journal of Medicinal Chemistry, 29(5), 641–648.

Ritchie, J.W. and P.M. Taylor.  (2001)  Role of the System L permease LAT1 in amino acid and iodothyronine transport in placenta.  Biochem. J.  356(Part 3); 719-725.

Riu, A., J.P. Cravedi, L. Debrauwer, A. Garcia, C. Canlet, I. Jouanin and D. Zalko. (2008) Environ. Int. 34(3): 318-329.

Roberts LM, Woodford K, Zhou M, Black DS, Haggerty JE, Tate EH, Grindstaff KK, Mengesha W, Raman C, Zerangue N 2008 Expression of the thyroid hormone transporters MCT8 (SLC16A2) and OATP14 (SLCO1C1) at the blood-brain barrier. Endocrinology 149:6251-6261

Rotroff, D.M., B.A. Wetmore, D.J. Dix, S.S. Ferguson, H.J. Clewell, K.A. Houck, E.L. Lecluyse, M.E. Anersen, R.S. Judson, C.M. Smith, M.A. Sochaski, R.J. Kavlock, F. Boellmann, M.T. Martin, D.M. Reif, J.F. Wambaugh and R.S. Thomas. (2010) Incorporating human dosimetry and exposure into high-throughput in vitro toxicity screening.  117(2): 348-358.

Sato, K., J. Sugawara, T. Sato, H. Mizutamari, T. Suzuki, A. Ito, T. Mikkaichi, T. Onogawa, M. Tanemoto, M. Unno, T. Abe and K. Okamura.  (2003)  Expression of organic anion transporting polypeptide E (OATP-E) in human placenta.  Placenta  24(2-3): 144-148.

Schreiber, G. (2002). The evolutionary and integrative roles of transthyrein in thyroid hormone homeostasis. Journal of Endocrinology, 175(1), 61–73. http://doi.org/10.1677/joe.0.1750061

Schroder van der Elst, J.P., D. van der Heide, H. Rokos, G. Morreale de Escobar and J. Kohrlre. (1998) Synthetic flavonoids cross the placenta in the rat and are found in fetal brain.  Am. J. Physiol.  274(2 Psrt 1): E253-E256.

Schroder van der Elst, J.P., D. van der Heide, H. Rokos, J. Kohrle and G. Morreale de Escobar. (1997)  Different tissue distribution, elimination, and kinetics of thyroxine and its conformational analog, the synthetic flavonoid EMD 49209 in the rat.  Endocrinology  138(1): 79-84.

Schuur, A.G., F.M. Boekhorst, A. Brouwer and T.J. Visser. (1997) Extrathyroidal effects of 2,3,7,8-tetrachlorodibenzo-p-dioxin on thyroid hormone turnover in male Sprague-Dawley rats.  Endocrinology  138(9): 3727-3734.

Sinjari, T. and P.O. Darnerud. (1998) Hydroxylated polychlorinated biphenyls: placental transfer and effects on thyroxine in the foetal mouse.  Xenobiotica  28(1): 21-30.

Sparkes, R.S., H. Sasaki, T. Mohandas, K. Yoshioka, I. Kilsak, Y. Sasaki, C. Heinzmann and M.I. Simon. (1987) Assignment of the prealbumin (PALB) gene (familial amyloidotic polyneuropathy) to human chromosome region 18q11.2-q12.1. Hum. Genet.  75(2): 151-154.

Stapleton, H.M., S.M. Kelly, R. Pei, R.J. Letcher and C. Gunsch.  (2009) Metabolism of polybrominated diphenyl ethers (PBDEs) by human hepatocytes in vitro.  Environ. Health Perspect.  117(2): 197-202.

Tohyama K, Kusuhara H, Sugiyama Y 2004 Involvement of multispecific organic anion transporter, Oatp14 (Slc21a14), in the transport of thyroxine across the blood-brain barrier. Endocrinology

Ucan-Marin, F., A. Arukwe, A.S. Mortensen, G.W. Gabrielsen and R.J. Letcher. (2010) Recombinant albumin and transthyretin transport proteins from two gull species and human: chlorinated and brominated contaminant binding and thyroid hormones.  Environ. Sci. Technol.  44(1): 497-504.

Van Birgelen, A.P., E.A. Smit, I.M. Kampen, C.N. Groeneveld, K.M. Case, J. Van der Kolk, H. Poiger, M. Van den Berg, J.H. Koeman and A. Brouwer. (1995) Subchronic effects of 2,3,7,8-TCDD or PCBs on thyroid hormone metabolism: use in risk assessment.  Eur. J. Pharmacol.  293(1) : 77-85.

Van den Berg, K. J. (1990). Interaction of chlorinated phenols with thyroxine binding sites of human transthyretin, albumin and thyroid binding globulin. Chemico-Biological Interactions, 76(1), 63–75.

Van den Berg, K. J., Van Raaij, J. a G. M., Bragt, P. C., & Notten, W. R. F. (1991). Interactions of halogenated industrial chemicals with transthyretin and effects on thyroid hormone levels in vivo. Archives of Toxicology, 65(1), 15–19.

Viberg, H., A. Fredriksson and P. Eriksson. (2002) Neonatal exposure to the brominated flame retardant 2,2',4,4',5-pentabromodiphenyl ether causes altered susceptibility in the cholinergic transmitter system in the adult mouse.  Toxicol. Sci. 67(1): 104-107.

Viollon-Abadie, C., D. Lassere, E. Debruyne, L. Nicod, N. Carmichael and L. Richert. (1999) Phenobarbital, beta-naphthoflavone, clofibrate, and pregnenolone-16alpha-carbonitrile do not affect hepatic thyroid hormone UDP-glucuronosyl transferase activity, and thyroid gland function in mice.  Toxicol. Appl. Pharmacol.  155(1) 1-12.

Visser, T.J. and R.P. Peeters. (2012) Metabolism of thyroid hormone.  In: De Groot LJ, Chrousos G, Dungan K, Feingold KR, Grossman A, Hershman JM, Koch C, Korbonits M, McLachlan R, New M, Purnell J, Rebar R, Singer F, Vinik A, editors. Endotext [Internet]. South Dartmouth (MA): MDText.com, Inc.; 2000-.

Visser, T. J. (2010). Cellular Uptake of Thyroid Hormones. In: De Groot LJ, Chrousos G, Dungan K, Feingold KR, Grossman A, Hershman JM, Koch C, Korbonits M, McLachlan R, New M, Purnell J, Rebar R, Singer F, Vinik A, editors. Endotext [Internet]. South Dartmouth (MA): MDText.com, Inc.; 2000-.

Visser, T.J. (1996) Role of sulfate in thyroid hormone sulfation.  Eur. J. Endocrinol.  134(1): 12-14.

Visser, T.J., E. Kaptein, J.A. van Raaij, C.T. Joe, T. Ebner and B. Burchell. (1993)

Multiple UDP-glucuronyltransferases for the glucuronidation of thyroid hormone with preference for 3,3',5'-triiodothyronine (reverse T3).  FEBS Lett.  315(1): 65-68.

Weiss, J.M., P.L. Andersson, M.H. Lamoree, P.E. Leonards, S.P. van Leeuwen and T. Hamers. (2009) Competitive binding of poly- and perfluorinated compounds to the thyroid hormone transport protein transthyretin.  Toxicol. Sci.  109(2): 206-216.

Weiss, J. M., Andersson, P. L., Zhang, J., Simon, E., Leonards, P. E. G., Hamers, T., & Lamoree, M. H. (2015). Tracing thyroid hormone-disrupting compounds: database compilation and structure-activity evaluation for an effect-directed analysis of sediment. Analytical and Bioanalytical Chemistry, 5625–5634. http://doi.org/10.1007/s00216-015-8736-9

Yamauchi, K., A. Ishihara, H. Fukazawa and Y. Terao.  (2003) Competitive interactions of chlorinated phenol compounds with 3,3',5-triiodothyronine binding to transthyretin: detection of possible thyroid-disrupting chemicals in environmental waste water.  Toxicol. Appl. Pharmacol.  187(2): 110-117.

Yen, P. M. (2001). Physiological and molecular basis of thyroid hormone action. Physiological Reviews, 81(3), 1097–1142.

Zhang, J., J.H. Kamstra, M. Ghorbanzadeh, J.M. Weiss, T. Hamers and P.L. Andersson. (2015) In Silico Approach To Identify Potential Thyroid Hormone Disruptors among Currently Known Dust Contaminants and Their Metabolites.  Environ. Sci. Technol.  49(16): 10099-10107.

Zoeller, R. T., Tan, S. W., & Tyl, R. W. (2007). General background on the hypothalamic-pituitary-thyroid (HPT) axis. Critical Reviews in Toxicology, 37(1-2), 11–53.

Zoeller, R.T. and J. Rovet.  (2004)  Timing of thyroid hormone action in the developing brain: clinical observations and experimental findings.  J. Neuroendocrinol.  16(10): 809-818.