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: 961

Key Event Title

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

Increased, Clearance of thyroxine from serum

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, Clearance of thyroxine from serum
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
Tissue

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
Organ term
liver

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
metabolic process 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
Nuclear receptor induced TH Catabolism and Developmental Hearing Loss KeyEvent Katie Paul Friedman (send email) Open for adoption Under Development
Hepatic nuclear receptor activation alters metamorphosis KeyEvent Jonathan Haselman (send email) Under Development: Contributions and Comments Welcome
TH displacement from serum TTR leading to altered amphibian metamorphosis KeyEvent Jonathan Haselman (send email) Under development: Not open for comment. Do not cite
TH displacement from serum TBG leading to altered amphibian metamorphosis KeyEvent Jonathan Haselman (send email) Under development: Not open for comment. Do not cite
thyroid follicular cell adenomas and carcinomas KeyEvent Charles Wood (send email) Under Development: Contributions and Comments Welcome
AhR activation in the liver leading to Adverse Neurodevelopmental Outcomes in Mammals KeyEvent Prakash Patel (send email) Under development: Not open for comment. Do not cite

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
Term Scientific Term Evidence Link
African clawed frog Xenopus laevis NCBI
Rattus sp. Rattus sp. NCBI
mouse Mus musculus NCBI

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

Thyroxin (T4) and T3 are metabolized and cleared from tissues in a number of ways: inner ring or outer ring deiodination via specific enzymes, conjugation (glucuronidation or sulfation), oxidative deamination and ether-linked cleavage (Zoeller et al 2007).

Deiodination:

There are three types of deiodinase enzymes. D1 and D2 convert T4 to T3 by removing an iodine atom from the outer ring while D3 removes an iodine atom from the inner ring, converting T4 to reverse T3. Differential expression of these enzymes during brain development are critical to the functionality of thyroid hormone in different areas of the fetal brain.

Much of the T4 is carried to the liver, where it is transported across the cellular membrane, converted into T3 via deiodination as mediated by deiodinase enzymes, and it is this T3 that triggers the TH receptors found in the nucleus. Roughly 80% of the T3 needed is produced via outer-ring deiodination of T4, which "activates" T4 to T3 (as opposed to inner-ring deiodination, which "degrades" T4 to reverse T3 which is eliminated). About 30% of the T4 produced daily (~ 130 nmol) is converted to roughly 40 nmol of T3 (Visser 2012) via enzyme D1 (liver, kidney) while conversion to rT3 accounts for roughly 40% of T4 turnover and is mediated via enzyme D3 (brain, placenta, fetus).

Conjugation:

Glucuronidation and sulfation of T4 accounts for the rest of the metabolized T4 and leads to rapid elimination through bile. It is thought that 20% of daily T4 production is eliminated through biliary excretion of glucuronide conjugates. Glucuronidation is carried out by UDP-glucuronoyltransferase (UGT) enzymes (Hood and Klaassen 2000a, 2000b) and appears to be more important in murine species than in man (Henneman and Visser 1997) and sulfation of T4 is done largely through an initial inner ring deiodination step (via D3). Circulating levels of THs in serum can be affected by compounds that induce the activity of UDP-UGT enzymes.

Uptake into the liver involves "high affinity, low capacity" and "low affinity, high capacity" processes with Km values in the nano- to micro-molar range (as opposed to the free T3 and T4 concentrations, which are in the picomolar range) (Henneman et al 2001 from Visser 2010). Both MCT8 and MCT 10 can transport THs; however, MCT8 is expressed in human liver where MCT10 is not and MCT8 display higher efficacy of cellular uptake and efflux relative to T3 (Ref 12 in Visser 2010).

Data in animals for PCBs

Previous reports have been made showing serum TH decreases in rats and mice in response to PCBs, PCB congeners and TCDD and these decreases have been thought to be driven by UDP-UGT (particularly 1A1 and 1A6)(Barter and Klaassen 1994, Schuur et al 1997, Van Birgelen et al 1995, Visser 1996)

Hallgren et al 2001, Hallgren and Darnerud 2002 showed that both T4 and T3 are significantly decreased following exposure to PCB mixtures or individual congeners.

Kato et al 2003 (and Kato et al 2002) showed for the first time that a commercial PCB mixture (Kanechlor-500, KC500) decreased serum TH without an increase in glucuronidation of T4. Male Wistar rats and ddy mice were given a single ip injection of 100 mg/kg and 4 days later, organ weights were measured and microsomal enzymes measured.  Significant increases were noted for both endpoints in both species; however, treatment with PCBs led to significant increases in UDP-UGT activity in rats but not mice. Gene expression of UDP-UGTs was also examined and, again, rats (but not mice) displayed time-dependent increases in levels of UGT1A1 and UGT1A6 following treatment with PCBs.  This agrees with past reports showing that clofibrate, phenobarbital, pregnenolone-16-alpha-carbonitrile and beta-naphthoflavone decrease serum TH and increase hepatic UDP-UGT activity in rats but not mice (Viollon-Abadie et al 1999).  

This implies that mice may reduce serum TH through a mechanism that does not involve increased glucuronidation, but may involve a TTR-associated pathway (as hydroxylated metabolites of PCBs have been displayed high affinity for TTR in in vitro studies).  Kanechlor-500 does not display any appreciable amount of outer ring deiodination activity (which would convert serum T4 to T3) and treatment with the mixture did not significantly change TSH levels (indicating there is no induction of the thyroid feedback loop from the measured decreases in serum TH). 

Kato et al 2004 performed the same experiment next with Wistar and Gunn rats, the latter species being a Wistar mutant strain that lacks UGT1A isoforms.  Both species showed serum T4 (free and total) decrease after single injection of either KC500 or pentaCB and only Wistar rats showed an associated increase in UDP-UGT activity. Significant decrease in type I deiodinase was observed in both rats in addition to detection of hydroxylated PCB metabolites bound to TTR.  Gunn rats treated with clofibrate also showed decreased serum T4 without an increase in UDP-UGT activity (Visser et al 1993).  These results imply that decreases of serum TH by PCB or pentaCB were managed by formation of OH-PCB metabolites that were then transported by TTR.  This is supported by the fact that the main metabolite found in KC500-treated rats was 4-OH-2,3,3’,4’,5-pentachlorobiphenyl, which displays a binding affinity towards TTR that exceeds that of T4 by more than 3-fold (Meerts et al 2002).  In fact, the dihydrohxylated PCBs show several fold higher affinity than the monohydroxylated PCBs (Lans et al 1993).  It should be noted that an increase in sulfation via SULT enzymes may also offer an esxplanation for the observed results.

Kato et al 2007 treated Wistar and Gunn rats with KC500 at a lower dose (10 mg/kg) once daily for 10 days, noting decrease in total and free serum T4 as well as the differential UDP-UGT response across the different strains.  Clearance of [125I]T4 from serum was higher in both species treated with KC500 and accumulation in several tissues, particularly the liver, was observed.  These data imply that reduction of serum TH from exposure to KC500 would be mediated through accumulation in the tissues and not through an increase in glucuronidation.  In addition, competitive inhibition by PCB or its metabolites with serum transport proteins (like TTR) could also decrease serum T4 by inducing a change in tissue distribution, especially the liver where more than 40% of[125I]T4 accumulated following treatment.

Kato et al 2009 treated C57BL/6 and DBA/2 mice with the heptaCB metabolite 4-OH-CB187, decreasing free and total serum T4 with no observed UDP-UGT activity or effect on TSH.  A number of OH-PCBs have been identified in human serum, including 4-OH-CB107, 3-OH-CB153, 4-OH-CB146, 3’-OH-CB138 and 4-OH-CB187 (which specifically has a 5-fold higher affinity for TTR relative to T4)(Hovander et al 2002).  Levels of [125I]T4-TTR were decreased with accompanying increases in binding to TBG and albumin in both strains of mice. Finally, T4 levels increased in tissues, particularly the liver and kidney.  Decreases in total and free serum T4 mediated by 4-OH-CB187 were observed in wild-type and TTR-heterozygous mice but not in TTR-deficient mice, with heterozygous mice displaying a smaller decrease in T4 relative to TTR-deficient mice.  In both strains of mice, treatment with 4-OH-CB187 promoted clearance of [125I]T4 from serum relative to controls and serum pharmacokinetic data were estimated, along with tissue-to-serum (Kp value) concentration ratios and [125I]T4 tissue distribution levels. These data imply that 4-OH-CB187 inhibits formation of the [125I]T4-TTR complex, which may lead to a change in tissue distribution, with accumulation in the liver and kidney mainly.

Kato et al 2012 treated C57BL/6 (wild type) and TTR-null mice with single ip injections of pentaCB at 112 mg/kg, noting significant decreases in total serum T4 and T4-TTR complex and measuring [125I]T4 clearance from serum and accumulation in tissues.  Treatment with pentaCB resulted in decrease of [125I]T4-TTR and increase in [125I]T4-albumin and [125I]T4-TBG complexes in wild type mice, but not in TTR-deficient mice, although liver accumulation was noted in both strains independent of UDP-UGT activity.  These data imply that penta-CB mediated increases in T4 liver concentration occurs mainly through inhibition of efflux of T4 and/or promotion on influx of T4 into hepatic cells (which is a receptor mediated process independent of TTR transport at the liver).

Kato et al 2013 treated C57BL/6 and DBA/2 mice with 50 mg/kg CB118 (pentaCB) in a single ip injection for 5 days, noting decreased serum T4 in both strains and decrease in TSH for the DBA/2 mice but not C57BL/6.   CB118-mediated changes in [125I]T4 complexes with TBG, albumin and TTR were only observed in C57BL/6 mice (and not DBA/2), despite [125I]T4 accumulation in the liver of both strains. It is thought that the strain differences are dependent on differences in induction of CYP1A enzymes responsible for the hydroxylation of PCBs (creating metabolites that display far greater affinity for TTR than the natural T4 ligand).

Martin and Klaassen 2010 treated male Sprague Dawley rats with Aroclors 1242 and 1254; PCBs 95, 99, 118, 126 or TCDD at 4 doses via gavage daily for 7 days, then measured serum TH via radioimmunoassay and induction of hepatic Cyp1a and Cyp2b.  This study was the first to examine all three classes of PCB congeners: TCDD-type (no chlorine substitutions in ortho position, high affinity for arylhydrocarbon receptor, induce Cyp1a, PCBs 77 and 126), PB-type (at least 2 ortho substitutions, low affinity for AhR, induce Cyp2b, PCBs 28, 95, 99, 101 and 153) or mixed type (1 ortho substitution, low affinity for AhR, induce both Cyp1a and Cyp2b, Aroclors and PCB 118).  This study showed that PB-type and mixed type PCB congeners are more effective than TCDD type in reducing serum T4, with Aroclor 1254 (mixed) and PCBs 99 (PB) and 118 (mixed) producing the greatest reduction in serum T4 (as well as T3).  Serum TSH was not affected by any compound.  Total and free serum T4 was decreased by all treatments in a dose-dependent manner; however marked reduction were noted following treatment with Aroclor 1254, PCB 99 and PCB 118. PCB 118 and 126 caused significant increase in Cyp1a activity while Aroclor 1254 and PCBs 99 and 118 significantly induced Cyp2b.  Thus, it appears TCDD type congeners induce CYP1A2 (EROD) activity and UGT-UDP activity in the liver (associated wth binding at AhR) while PB type congeners induce CYP1B2 (PROD) activity and do not induce UGT-UDPs in the liver (associated with increased tissue uptake).

The PB type congeners may induce Oatp1a4 activity to increase clearance from plasma and enhance tissue uptake. Guo et al 2002 reported increase of Oatp1a4 following treatment with PCB 99 (and a decrease following treatment with PCB 126, a TCDD type congener). There are also reports of PB type congeners that accumulate in the liver with little to no increase in glucuronidation or biliary excretion and no changes in serum binding proteins, such as PCB 153, which implies a possible induction of OATP hepatic cellular transport proteins (Kato et al 2011).

Martin et al 2012 treated male Wistar rats with Aroclors 1242 and 1254, PCBs 95, 99, 118 and 126 and TCDD via gavage one per day for 7 days, followed 24 hours later with injection of [125I]T4 and collection of urine, blood, bile and urine. No treatments increased urinary excretion of [125I]T4, but serum T4 was reduced in all treatments and biliary excretion increased following treatment of Aroclor 1254, PCBs 118 and 126, and TCDD as measured by induction of UDP-UGT activity in the liver. PCBs 95 and 99 (PB type congeners) did not induce UGT-UDP activity despite very large and rapid decrease of serum [125I]T4 by PCB 99.  These data imply that increased tissue uptake (perhaps through increased TH transport across cell membranes) is another mechanism by which serum T4 can be reduced. 

Kato et al 2013 showed that PCB 118 (mixed type) mediated changes in tissue distribution and transport proteins in C57BL/6 mice, but not DBA/2 mice.  Kato et al 2012 showed the same with synthesized 2,2’,4,5,5’-pentaCB (PCB 101, PB type). Kato et al 2014 showed that PCB 77 (TCDD type) mediated changes in tissue distribution and transport proteins in DBA/2 mice, but not C57BL/6 mice.

Erratico et al 2012 used pooled and single-donor human liver microsomes, human recombinant cytochrome P450 (CYP) enzymes and CYP-specific antibodies to evaluate the oxidative metabolism of BDE-99.  Ten (10) hydroxylated metabolites were produced by human microsomes and identified via HPLC-MS/MS and rates of formation were determined, including several that are much more potent than the natural ligand.  All ten were found to be catalyzed solely by CYP2B6.  Previous studies had also shown formation of hydroxylated metabolites of BDE-99 by human hepatic preparations (Lupton et al 2009, 2010; Stapleton et al 2009); however, fewer OH-PBDEs and additional CYP enzymes were found in similar work done with rat microsomes (Erratico et al 2011).

Feo et al 2013 incubated BDE-47 and recombinant CYPs, measuring the metabolites via GC-MS/MS, as well as specific kinetic studies with BDE-47, CYP2B6 and pooled human liver microsomes.  Six (6) OH-PBDEs were found to be catalyzed by CYP2B6 and additional metabolites were identified upon GC-MS/MS (including the novel finding of dihydroxylated metabolites) and these metabolites have been previously found in human serum (Athanasiadou et al 2008; Qui et al 2009).  The kinetic studies showed that hydroxylation can occur at low concentrations and that CYPT2B6 has high affinity for BDE-47.  CYP2C19 and CYP3A4 were also suggested to play minor roles in the formation of OH-PBDEs.

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

Methods that have been previously reviewed and approved by a recognized authority should be included in the Overview section above. All other methods, including those well established in the published literature, should be described here. Consider the following criteria when describing each method: 1. Is the assay fit for purpose? 2. Is the assay directly or indirectly (i.e. a surrogate) related to a key event relevant to the final adverse effect in question? 3. Is the assay repeatable? 4. Is the assay reproducible?

Thyroid hormone uptake into human tissues has been measured by analyzing the rate of disappearance of radiolabeled TH from plasma into rapidly and slowly equilibrating tissue compartments (Visser 2010).

Measuring the rate of T4 glucuronidation and sulfation as well as biliary excretion informs the mechanism of action of thyroid system modulation. Studies involving knock/out mice and thyroidectomized rats also inform this mechanism.

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.