ArticlePDF Available

Effect of Altered Thyroid Status on Neurotrophin Gene Expression During Postnatal Development of the Mouse Cerebellum

Authors:

Abstract and Figures

Thyroid hormone (TH) plays an important role in brain development. The rodent cerebellum has been an excellent model for the study of the molecular mechanisms of TH action in brain. However, most studies have utilized the rat rather than the mouse. Considering the usefulness of mice with regard to diverse genetic models, the study of TH effect on mouse cerebellar development is needed. Thus, we examined the effect of perinatal hypothyroidism on the expression of neurotrophin-3 (NT-3) and brain-derived neurotrophic factor (BDNF) genes, which play critical roles in cerebellar development. Newborn mice were rendered hypothyroid by administering methimazole and perchlorate in drinking water to their mothers. The growth of hypothyroid mice was retarded, which was reversed by daily thyroxine administration. NT-3 and BDNF gene expression was depressed in the perinatal hypothyroid cerebellum. Furthermore, the expression of retinoid-receptor-related orphan nuclear hormone receptor-alpha (RORalpha), an orphan nuclear receptor that plays critical roles in Purkinje cell development, was also decreased. Morphologically, disappearance of the external granule cell layer was retarded and arborization of Purkinje cell dendrite was decreased, events that were also observed in hypothyroid rats. These results indicate that the mouse cerebellum is comparable to the rat cerebellum as a model for the examination of the molecular mechanisms of TH action in brain development.
Content may be subject to copyright.
American Thyroid Association Guide to Investigating
Thyroid Hormone Economy and Action
in Rodent and Cell Models
Report of the American Thyroid Association Task Force
on Approaches and Strategies to Investigate Thyroid Hormone Economy and Action
Antonio C. Bianco,
1,
*
Grant Anderson,
2
Douglas Forrest,
3
Valerie Anne Galton,
4
Bala´ zs Gereben,
5
Brian W. Kim,
1
Peter A. Kopp,
6
Xiao Hui Liao,
7
Maria Jesus Obregon,
8
Robin P. Peeters,
9
Samuel Refetoff,
7
David S. Sharlin,
10
Warner S. Simonides,
11
Roy E. Weiss,
7
and Graham R. Williams
12
Background: An in-depth understanding of the fundamental principles that regulate thyroid hormone homeostasis
is critical for the development of new diagnostic and treatment approaches for patients with thyroid disease.
Summary: Important clinical practices in use today for the treatment of patients with hypothyroidism, hyper-
thyroidism, or thyroid cancer are the result of laboratory discoveries made by scientists investigating the most
basic aspects of thyroid structure and molecular biology. In this document, a panel of experts commissioned by the
American Thyroid Association makes a series of recommendations related to the study of thyroid hormone
economy and action. These recommendations are intended to promote standardization of study design, which
should in turn increase the comparability and reproducibility of experimental ndings.
Conclusions: It is expected that adherence to these recommendations by investigators in the field will facilitate
progress towards a better understanding of the thyroid gland and thyroid hormone dependent processes.
INTRODUCTION
O
ver the past 150 years, investigators utilizing animal
and cell culture–based experimental models have
achieved landmark discoveries that have shaped our under-
standing of thyroid physiology and disease. From the iden-
tification of the long-acting thyroid stimulator to the
discovery of antithyroid drugs, basic research studies have
provided the fundamentals upon which our clinical diag-
nostic and therapeutic tools are based. Tens of thousands of
publications indexed on PubMed (www.pubmed.gov) fea-
ture cells or small animals made hypothyroid or thyrotoxic.
The great similarities in multiple aspects of thyroid physi-
ology between humans and small rodents have facilitated the
rapid translation of experimental findings to the clinical
realm. At the same time, fundamental interspecies differences
do exist and must be carefully accounted for if the experi-
mental findings are to have clinical relevance.
1
Division of Endocrinology, Diabetes and Metabolism, University of Miami Miller School of Medicine, Miami, Florida.
2
Department of Pharmacy Practice and Pharmaceutical Sciences, College of Pharmacy, University of Minnesota Duluth, Duluth, Minnesota.
3
Laboratory of Endocrinology and Receptor Biology, National Institute of Diabetes and Digestive and Kidney Diseases, National Institutes of
Health, Bethesda, Maryland.
4
Department of Physiology and Neurobiology, Dartmouth Medical School, Lebanon, New Hampshire.
5
Department of Endocrine Neurobiology, Institute of Experimental Medicine, Hungarian Academy of Sciences, Budapest, Hungary.
6
Division of Endocrinology, Metabolism, and Molecular Medicine, and Center for Genetic Medicine, Feinberg School of Medicine,
Northwestern University, Chicago, Illinois.
7
Section of Adult and Pediatric Endocrinology, Diabetes, and Metabolism, The University of Chicago, Chicago, Illinois.
8
Institute of Biomedical Investigation (IIB), Spanish National Research Council (CSIC) and Autonomous University of Madrid, Madrid,
Spain.
9
Division of Endocrinology, Department of Internal Medicine, Erasmus Medical Center, Rotterdam, The Netherlands.
10
Department of Biological Sciences, Minnesota State University, Mankato, Minnesota.
11
Laboratory for Physiology, Institute for Cardiovascular Research, VU University Medical Center, Amsterdam, The Netherlands.
12
Department of Medicine, Imperial College London, Hammersmith Campus, London, United Kingdom.
*Chair; all other authors are listed in alphabetical order.
THYROID
Volume 24, Number 1, 2014
ª American Thyroid Association
DOI: 10.1089/thy.2013.0109
88
While certain experimental techniques have been widely
accepted and adapted following their use in papers gener-
ated by influential labs, lack of standardization has un-
doubtedly promoted heterogeneity of results. Because
certain experimental variables may have unknown biologi-
cal threshold levels, lack of s tandardization may lead to have
highly discordant results in different studies examining the
same issue.
To address this lack of standardization, the American
Thyroid Association (ATA) convened a panel of specialists in
the field of basic thyroid research to define consensus strate-
gies and approaches for thyroid studies in rodents and in cell
models. This task force was charged with reviewing the lit-
erature first to determine which experimental practices could
benefit from standardization and second to identify critical
experimental variables that demand consideration when
thyroid studies are being designed. The conclusions of the
task force are presented in this document as ‘American
Thyroid Association Guide to Investigating Thyroid Hor-
mone Economy and Action in Rodent and Cell Models.’ The
70 recommendations and their accompanying commentaries
examine topics ranging from ‘making cells hypothyroid’ to
‘how to study the thyrotoxic bone.’ While far from exhaus-
tive, these recommendations touch on certain fundamental
aspects of thyroid research relevant for all investigators in
the field.
Each recommendation in thi s guide promotes a particular
experimental approach based o n criteria including th e
prevalence of the approach, with widely used techniques
being given precedence, a nd in particular whether the ap-
proach has been shown to lead to reproducible results in
studies by independent investigators. Because head-to-head
scientific comparisons of experimental methods in this field
are virtually nonexistent, these recommendations cannot be
graded on the basis of strength of evidence in the fashion o f
clinical guidelines; indeed, all would be graded as ‘expert
opinion.’ At the same tim e, unli ke clin ical guid elines, the
main goal of these recommendations and their accompany-
ing commentaries is not to identify the single best practice
per se, but instead to encourage investigators to choose
standard approaches; f or example, avoiding r andom treat-
ment doses or methods of thyroid hormone administration,
which would only serve to limit comparison with previous
studies.
The practical nature of recommendations should become
readily apparen t to the reader. This document is inte nded
to serve as a reference for investigators, assisting them in
making design choices that avoid well-known pitfalls
while increasing standardization in the field. As part o f
this practical approach, reference credit is often given to
manuscripts in which the technical details are most clearly
or comprehensively explained, rather t han the first publi-
cation to use a technique. In addition, emphasis was
placed on contemporary approaches, rather than historical
strategies, such that the document illustrates what is cur-
rently available for the contemporary study of thyroid
hormone homeostasi s, metabolism, and action . It is the
position of the ATA that animal studies should b e per-
formed in accordance with all applicable ethical s tandards
and research protocols approved by local institutional
animal committees.
METHODS OF DEVELOPMENT
OF RECOMMENDATIONS
Administration
The ATA Executive Council selected a chairperson to lead the
task force, and this individual (A.C.B.) identified the other 14
members of the panel in consultation with the ATA board of
directors. Membership on the panel was based on expertise and
previous contributions to the thyroid field. Panel members de-
clared whether they had any potential conflict of interest during
the course of deliberations. Funding for the guide was derived
from the ATA and thus the task force functioned without
commercial support.
To develop a useful document, the task force first devel-
oped a list of the topics that would be most helpful and the
most important questions that scientists working in the thy-
roid field might pose when planning an experiment or inter-
preting experimental data. Each of the 10 topics was
distributed to a primary writer who used his or her knowl-
edge of the subject as well as a systematic PubMed and
Google Scholar search for primary references, reviews, and
other materials publicly available before December 2012, to
develop a set of recommendations. All drafts were reviewed
and edited by the chair for consistency and sent back to the
primary writers for review; in some cases multiple iterations
took place until the recommendation was finalized. A pre-
liminary draft of each recommendation was then reviewed by
secondary and tertiary reviewers within the group who then
prepared additional critiques. These were addressed by the
primary writer and sent back to the chair. All drafts were
merged and posted at a protected web address available only
to the task force members and ATA office. This document
remained available for periodic review by the task force at
large, with critiques and suggestions sent back to the chair
that updated the document. In a few cases the chair asked for
outside experts to critically review specific recommendations
given their expertise in a focused area. Their comments and
suggestions were then worked into the master document, and
the contributions are acknowledged at the end of this article.
The panel agreed that recommendations would be based
on consensus of the panel. Task force deliberations took place
largely through electronic communication. There were also a
few meetings of the authors and telephone conference calls.
Presentation, Approval, and Endorsement
of Recommendations
The structure of our recommendations is presented in Table
1. Specific recommendations are presented within the main
body of the text and in many cases broken down in subitems
identified by letters. The page numbers and the location
key can be used to quickly navigate to specific topics and
recommendations.
Prior to the initial submission of these guidelines, they were
approved by the board and executive committee of the ATA
and afterwards submitted to the membership of the ATA in
early 2013 for comments and suggestions. This feedback was
considered in the further preparation of the document that
was submitted for publication. Subsequent to the document
being accepted for publication in Thyroid, it was approved by
the board and executive committee of the ATA.
INVESTIGATING THYROID HORMONE ECONOMY AND ACTION 89
Table 1. Organization of the Task Force’s Recommendations
Location
key Sections and subsections Page
Location
key Sections and subsections Page
T
3
, 3,3¢,5-triiodothyronine; TR, thyroid hormone receptor; PCR, polymerase chain reaction.
[A] Assessing the Thyroid Gland 91
[A.1] Structure–function relationships 91
Recommendation 1 91
[A.2] Thyroid iodide kinetics 93
Recommendation 2 94
Recommendation 3 95
[A.3] Thyroid imaging 95
Recommendation 4 95
[B] Assessing Circulating and Tissue
Thyroid Hormone Levels
97
[B.1] Serum 97
Recommendation 5 98
Recommendation 6 99
Recommendation 7 100
[B.2] Tissue 100
Recommendation 8 100
[B.3] Sources of tissue T
3
and TR saturation 100
Recommendation 9 101
[C] Assessing Thyroid Hormone
Transport Into Cells
101
[C.1] Thyroid hormone transport
in vitro
102
Recommendation 10 102
Recommendation 11 103
[C.2] Thyroid hormone transport in vivo 103
Recommendation 12 103
[D] Assessing Thyroid Hormone Deiodination 104
[D.1] Identification, expression, and
quantification of deiodinases
104
Recommendation 13 104
Recommendation 14 105
[D.2] Deiodination in intact cells 106
Recommendation 15 106
[D.3] Deiodination in perfused organs 106
Recommendation 16 106
[D.4] Deiodination in whole animals 107
Recommendation 17 107
[D.5] Non-deiodination pathways of
thyroid hormone metabolism
108
Recommendation 18 109
[E] Inducing Hypothyroidism and
Thyroid Hormone Replacement
109
[E.1] Hypothyroidism in animals 109
Recommendation 19 109
Recommendation 20 110
Recommendation 21 111
Recommendation 22 111
Recommendation 23 112
[E.2] Thyroid hormone replacement
in animals
113
Recommendation 24 113
[E.3] Hypothyroidism in cultured cells 114
Recommendation 25 114
[F] Increasing Thyroid Hormone
Signaling
114
[F.1] Thyrotoxicosis in animals 114
Recommendation 26 115
Recommendation 27 115
[F.2] Thyrotoxicosis in cultured cells 115
Recommendation 28 115
[F.3] Use of thyroid hormone analogues 116
Recommendation 29 116
[G] Iodine Deficiency and Maternal–Fetal
Transfer of Thyroid Hormone
117
[G.1] Iodine deficiency in rodents 117
Recommendation 30 117
Recommendation 31 118
Recommendation 32 118
[G.2] Placental transfer of thyroid hormone 118
Recommendation 33 118
[H] Models of Nonthyroidal Illness 118
Recommendation 34 119
Recommendation 35 119
[I] Assessing Thyroid Hormone Signaling
at Tissue and Cellular Levels
119
[I.1] Gene expression as a marker of
thyroid hormone status
120
Recommendation 36 120
[I.2] PCR analysis of mRNA expression levels 120
Recommendation 37 120
[I.3] Genome-wide analysis of thyroid
hormone-responsive mRNA
122
Recommendation 38 122
[I.4] Mechanisms of gene regulation by
thyroid hormone
122
Recommendation 39 122
Recommendation 40 123
[I.5] Mouse models for indicating thyroid
hormone and TR signaling in tissues
123
Recommendation 41 124
[J] Assessing Thyroid Hormone Signaling
by Way of Systemic Biological Parameters
124
[J.1] Central nervous system 125
Recommendation 42 126
Recommendation 43 126
Recommendation 44 127
Recommendation 45 127
Recommendation 46 127
Recommendation 47 128
Recommendation 48 128
[J.2] Heart and cardiovascular system 129
Recommendation 49 129
Recommendation 50 129
Recommendation 51 130
Recommendation 52 130
Recommendation 53 131
Recommendation 54 132
[J.3] Intermediary metabolism and
energy homeostasis
132
Recommendation 55 132
Recommendation 56 135
Recommendation 57 135
Recommendation 58 136
Recommendation 59 137
[J.4] Skeletal muscle 137
Recommendation 60 138
Recommendation 61 138
Recommendation 62 138
Recommendation 63 139
Recommendation 64
139
Recommendation 65 139
[J.5] Skeleton 140
Recommendation 66 140
Recommendation 67 140
Recommendation 68 140
Recommendation 69 140
Recommendation 70 142
90
The final document was officially endorsed by the American
Academy of Otolaryngology–Head and Neck Surgery (AAO-
HNS), American Association of Endocrine Surgeons (AAES),
American College of Nuclear Medicine (ACNM), Asia and
Oceania Thyroid Association (AOTA), British Nuclear Medi-
cine Society (BNMS), British Thyroid Association (BTA),
European Thyroid Association (ETA), International Associa-
tion of Endocrine Surgeons (IAES), Italian Endocrine Society
(SIE), Japan Thyroid Association (JTA), Korean Society of
Head and Neck Surgery (KSHNS), Latin American Thyroid
Society (LATS), Korean Society of Nuclear Medicine (KSNM)
and The Endocrine Society (TES).
RESULTS
[A] Assessing the Thyroid Gland
Overview. Studies of function–structure relationship of
the thyroid gland, as well as studies of thyroid iodide kinetics
and imaging are traditionally employed to assess the thyroid
gland. Structural characterization is important to assess
functional changes such as hypo- and hyperthyroidism and
for evaluating transformation of thyroid cells into a malignant
phenotype (1–3). At the same time, the study of thyroidal
iodide economy and thyroid imaging are relevant not only to
studies of thyroid hormone synthesis but also to under-
standing the effects of environmental toxins such as perchlo-
rate or thiocyanate on thyroid economy (4–7).
[A.1] Structure–function relationships
Background. While the human thyroid consists of a left
and a right lobe that are connected by an isthmus, rodents
have two independent thyroid lobes. The thyroid gland is
divided by connective tissue septa into lobules, each one of
these containing from 20 to 40 follicles, the basic functional
unit of the thyroid gland. The follicle is a round or elongated
hollow structure lined by a single layer of polarized cuboidal
or flattened follicular cells that is filled with thyroglobulin-
containing colloid. It is surrounded by a basal membrane and
a rich capillary network with high blood flow (8). The follicles
normally vary considerable in size, and the follicular cell
morphology is usually monotonous. The height of the cells
varies according to the functional status of the gland.
&
RECOMMENDATION 1a
Morphometry of thyroid follicles can be used as an index of
thyroidal activity.
Commentary. The entire gland should always be dis-
sected while attached to the trachea and immediately fixed
with 10% neutral buffered formalin for histological and im-
munohistochemical analysis. Hematoxylin and eosin (H&E)
staining is widely used to assess the thyrocytes, whereas
periodic-acid Schiff staining stains thyroglobulin avidly and is
well suited to highlight follicular protein content and follic-
ular structure (Fig. 1) (8). Structural modifications reflect
changes in secretory activity resulting from iodine deficiency
(9), chronic cold exposure (10), or treatment with antithyroid
drugs (11). Some follicular cell parameters such as height can
be measured under light microscopy using an ocular mi-
crometer grid (e.g., in a 1-month-old rat, the epithelial cell
height is about 10 lm) (12). A flat epithelium is hypoactive,
while a heightened epithelium is observed in glands in which
the thyrotropin (TSH) pathway is stimulated (10). The use of
computerized semiautomatic image analysis is more objective
and used widely (13). Such morphometric analysis should be
focused on one of the central sections of the thyroid (13) that is
representative of the whole lobe (14). The data obtained are
reduced by predefined mathematical models that assume
thyroid follicles have a spherical shape and follicular cells are
octagons with a square base. This data reduction yields the
following parameters: mean follicle circumference; surface
area and volume; total volume of epithelium and colloid;
number of epithelial cell nuclei visible in each follicle; and the
height, surface area, and volume of thyroid epithelial cell,
which can be used to estimate the functional state of the
thyroid gland. Thus, the activation index, expressed by the
epithelial volume/colloid volume ratio, increases as the thy-
roid becomes more active, reflecting an increase in the epi-
thelial volume and a decrease in the colloid volume (13).
Measurement of total cell volume in cultures of primary
thyrocytes or cell lines cultured in vitro can be performed
using confocal laser-scanning microscopy after cells are loa-
ded with octadecylrhodamine B (15,16).
FIG. 1. Microscopic structure of the mouse thyroid. (A)
Hematoxylin and eosin (H&E) staining. (B) Periodic acid
Schiff (PAS) staining. Mice were euthanized, and the thy-
roids dissected, fixed in buffered formalin, and embedded in
paraffin. Thyroid sections (5 lm) were mounted on glass
slides, de-paraffinated, and hydrated. For histological anal-
ysis, sections were stained with H&E, following a standard
protocol. Glycoproteins were detected using PAS staining.
Sections were stained with 0.5% periodic acid for 30 minutes
and with Schiff’s reagent for 20 minutes and then rinsed in
running tap water for 5 minutes. Nuclei were counterstained
with hematoxylin for 3 minutes. Sections were rinsed in
running tap water, dehydrated, cleared, and mounted. Re-
produced with permission from Senou et al. (20).
INVESTIGATING THYROID HORMONE ECONOMY AND ACTION 91
&
RECOMMENDATION 1b
Autoradiography can be used to quantify the overall ac-
tivity of thyroid follicles and to determine the location of
iodide within follicles.
Commentary. Thyroid follicular cells concentrate iodide
according to their activity. Although the activity of the thou-
sands of follicular cells should be similar within a given thy-
roid gland, there is a great deal of variation among cells within
the same follicle and between follicles. Thus autoradiography
provides unique insights into the activity of individual thy-
roid follicular cells.
125
I is injected intravenously, typically 48 to 72 hours prior to
killing the animal. Thyroid glands are dissected and processed
for autoradiography using standard techniques (17,18). Orga-
nification of iodide can be blocked by treatment of the animals
with methimazole (MMI). Autoradiography experiments with
human, rodent, and feline goiter tissue have also been per-
formed after xenotransplantation of thyroid tissue into nude
mice. Subcutaneously implanted fragments are maintained in
recipient mice for several weeks before further analysis (19).
&
RECOMMENDATION 1c
The ultrastructural distribution of iodide within thyroid
follicles can be defined with secondary ion mass spec-
trometry (SIMS).
Commentary. SIMS is a technique used to analyze the
composition of thin films by sputtering the surface of the
specimen with a focused primary ion beam and collecting and
analyzing ejected secondary ions (Fig. 2). The mass/charge
ratios of these secondary ions are measured with a mass
spectrometer to determine the elemental, isotopic, or molec-
ular composition of the surface to a depth of 1–2 nm. SIMS is
the most sensitive surface analysis technique, with elemental
detection limits ranging from parts per million to parts per
billion. It is uniquely suited for the study of trace ions distri-
bution at the ultrastructural level (20).
Ionic images show that the early distribution of iodine is
heterogeneous from one follicle to another, from one thyr-
ocyte to another inside the same follicle, and that this dis-
tribution varies as a function of time (21). In normal thyroids
the natural
127
I isotope is found predominantly in the fol-
licular lumina. The identification of lumina devoid of
127
I
and/or the demonstration of significant amounts of
127
Iin
the cytoplasm of the epithelial cells or on the apical mem-
brane indicates impairment of the iodination pathway. To
define the ultrastructural distribution of iodide using SIMS,
thyroid lobes are processed in a similar way as for electron
microscopy, including fixation with glutaraldehyde and
preparation of semithin sections (20).
&
RECOMMENDATION 1d
Confocal microscopy in conjunction with immunohisto-
chemistry (IHC) can be used for two- or three-dimensional
(2D or 3D) image reconstruction to study protein expres-
sion in thyroid follicles, the surrounding capillary network,
and the stroma.
Commentary. Antibodies are available against most key
proteins in thyrocyte biology (22,23). Thus, standard IHC
techniques are commonly used in thyroid studies (Figs. 3 and
4) (24,25). Visualization can be performed with conventional
light microscopy, immunofluorescence microscopy, or con-
focal microscopy for higher resolution and 2D or 3D image
reconstruction (26). Cell surface proteins and processes are
best investigated using scanning electron microscopy (10).
Endogenous peroxidase activity is very high in thyroid
cells and is detected by reacting fixed tissue sections with 3,3¢-
diaminobenzidine substrate; pretreatment with hydrogen
FIG. 2. Mouse thyroid transmission electron microscopy. Thyroid lobes were fixed in 2.5% glutaraldehyde in 0.1 M cacodylate
buffer for 1.5 hours, post-fixed in 1% osmium tetroxide for 1 hour, and embedded in LX112 resin (Ladd Research Industries,
Burlington, VT). (A) Thin sections (0.5 lm) were stained with toluidine blue and analyzed for morphology by light microscopy.
(B) Ultrathin sections were prepared and stained with uranyl acetate and lead citrate and examined with an electron microscope
Zeiss EM169 (Carl Zeiss, Oberkochen, Germany). (C) Ultrastructural distribution of
127
I by secondary ion mass spectrometry
(SIMS) imaging. Semi-thin sections were prepared, and the ultrastructural distribution of the iodide natural isotope (
127
I) was
obtained through imaging by SIMS, using the NanoSIMS 50 system. Maps were acquired under standard analytic conditions: a
Cs+ primary beam with impact energy of 16 keV and a probe with current intensity of 1 pA. The analyzed surface was
30 · 30 lm. Under these conditions, a lateral resolution of 100 nm is expected. All images were acquired in 256 · 256 pixels with a
counting time of 20 milliseconds per pixel. White areas correspond to iodine detection.
127
I is homogeneously distributed in the
follicular lumina and in a few intracytoplasmic vesicles. Reproduced with permission from Senou et al. (20).
92 BIANCO ET AL.
peroxide prior to incubation with primary antibody eliminates
endogenous peroxidase activity that will interfere in IHC
studies. The use of fluorescent-tagged proteins should be avoi-
ded if autofluorescence is a problem (as assessed by viewing
tissue sections with a fluorescence microscope before any anti-
body incubation). Fine subcellular distribution studies can
be done with IHC and confocal microscopy; immunogold
staining electron microscopy allows detection of antigens at
very high resolution in studies of subcellular distribution (Fig. 5)
(20,27).
[A.2] Thyroid iodide kinetics
Background. The synthesis of thyroid hormone, its tetra-
iodinated form thyroxine (T
4
), and 3,3¢,5-triiodothyronine (T
3
)
requires a normally developed thyroid gland, an adequate io-
dide intake, and a series of regulated biochemical steps in
thyroid follicular cells, which form the spherical thyroid folli-
cles, the functional unit of the thyroid gland (28). In thyroid
epithelial cells, the sodium iodide symporter (NIS) mediates
the iodide uptake into thyroid follicular cells (29), and its ex-
pression is polarized (i.e., it is expressed only in the basolateral
membrane). At the basolateral membrane of thyrocytes, Na
+
/
K
+
-ATPase generates a sodium gradient that permits NIS to
mediate perchlorate inhibitable, Na
+
-dependent iodide uptake
(30). Iodide then translocates to the apical membrane and
reaches the follicular lumen through the apical membrane.
While it has been assumed that iodide moves across the apical
membrane primarily because of the electrochemical gradient,
studies in frozen section demonstrated that it is first accumu-
lated in the cytoplasm and only later in the lumen, and apical
iodide efflux is rapidly accelerated in polarized cells after ex-
posure to TSH (31). Electrophysiological studies using inverted
plasma membrane vesicles suggested the existence of two
apical iodide channels, but their molecular identity has not
been determined (32). The multifunctional anion exchanger
pendrin (SLC26A4/PDS), which has affinity for anions such as
iodide, chloride, and bicarbonate is thought to represent one of
these entities (33,34). Both NIS and SLC26A4 expression and
activity are increased by TSH (30, 33). While the term iodide
uptake can be used broadly for in vitro and in vivo approaches,
data interpretation should take into account the critical differ-
ences between the two settings, with the former reflecting cel-
lular iodide uptake and the latter mainly the concentration of
organified iodine in the colloid.
FIG. 4. Detection of dual oxidase
(DUOX) and thyroperoxidase in
the mouse thyroid by immunohis-
tochemistry. (A) DUOX was de-
tected on frozen sections with rabbit
polyclonal antibody diluted 1/3000
and incubated overnight. Positivity
is observed at the apical pole
(arrows, inset). (B) thyroperoxidase
was detected on paraffin sections
with rabbit antibody LoadTPO
821, 4 lg/mL and incubated for
3 hours. Reproduced with permis-
sion from Senou et al. (20).
FIG. 3. Detection of thyroglobulin and iodinated thyro-
globulin in the mouse thyroid by immunohistochemistry. (A)
Thyroglobulin was detected on paraffin sections using anti-
thyroglobulin rabbit polyclonal antibody (Dako) diluted 1/
1500 and incubated overnight. (B) Iodinated thyroglobulin
was detected using mouse monoclonal antibody (B1) diluted
1/3000 and incubated overnight. Negative controls included
the replacement of primary antibody by the preimmune
serum or absence of the primary antibody. Reproduced with
permission from Senou et al. (20).
INVESTIGATING THYROID HORMONE ECONOMY AND ACTION 93
&
RECOMMENDATION 2a
Basolateral cellular iodide uptake and apical efflux of iodide
can be studied in monolayers of polarized cells cultured on
semi permeable membranes forming a two-chamber system,
or in nonpolarized cell models such as the FRTL-5 or PCCL3
rat thyroid cell lines.
Commentary. Measurement of iodide uptake and efflux in
nonpolarized cells is relatively straightforward. The establish-
ment of polarized cell systems requires isolation of primary cells
or transfection or transduction of polarized heterologous cells
and the documentation of intact monolayer formation, which
are tedious and time-consuming (31). For the iodide uptake
assays, cells are incubated in an uptake solution typically con-
taining 10
- 5
MNa
125
I for a desired time period. Organification
can be blocked by treating the cells with MMI. The intracellular
iodide content is determined by measuring radiolabeled iodide
in the cell lysates using a gamma counter after cell lysis. Results
are expressed as counts per minute per well or, ideally, per
microgram of DNA. The gravimetric amount of intracellularly
accumulated iodide (pmol/lg DNA) can also be calculated
based on the specific activity of the tracer. Alternative methods
that have been used include the use of halide quenchers. A
problem with this approach is that these quenchers are not
specific for iodide, but also react to other halides. The avail-
ability of a modified enhanced yellow fluorescent protein
(EYFP) H148Q/I152L with high affinity for iodide has allowed
tracking iodide influx and efflux with relatively good accuracy
and a high degree of correlation with direct measurements of
radiolabeled iodide (35–37). Alternatively, mass spectrometry
has been used to study the uptake of perchlorate into FRTL-5
cells, which is also mediated by NIS (38).
A number of cell models and setups are available to study
NIS-mediated iodide transport (4–6,31,33,39–41), including
multiple heterologous cell lines transiently expressing NIS
(29,42). Such studies are useful for the characterization of NIS
function and the activities of naturally occurring or artificial
mutant proteins (29,42). For example, they are useful to
measure steady-state and initial rate iodide uptake as well as
kinetic parameters of NIS-mediated iodide transport. Uptake
of iodide has also been studied in cancer cell lines transfected
or transduced with constructs in which the NIS cDNA is
under the control of tissue-specific promoters with the aim to
promote uptake of
131
I and to induce cell death through its
beta-emission (43,44). For studies assessing the effect of TSH,
the medium used to culture thyroid cells is changed to TSH-
deprived media for several days and then submitted to the
different experimental conditions.
&
RECOMMENDATION 2b
Iodide efflux from thyrocytes can be assessed in perchlorate-
treated thyroid cell lines.
Commentary. To study iodide efflux in vitro, cells are
loaded with
125
I for 1–2 hours and subsequently treated with
perchlorate in order to block iodide uptake by NIS. The efflux
can then be studied by collecting supernatants at one or
multiple time points (5). The intracellular content of iodide
should also be determined at one or multiple time points.
Another strategy is to use a two-chamber system, in which the
efflux of iodide at the apical membrane can be measured by
collecting the supernatant at one or multiple time points
(31,40). Measuring iodide directly with ion-selective elec-
trodes in supernatants or cell lysates is problematic because
these probes are not specific for iodide and also recognize
other halides such as chloride.
Efflux of radioactive iodide by the anion channel SLC26A4
(pendrin) or any other anion channels can also be studied in
multiple heterologous cell lines transiently expressing NIS that
allows for initial iodide uptake (3,40). This can be documented
by measuring intracellular iodide content in cells co-expressing
NIS and the channel of interest with direct comparison to cells
that only express NIS. A model system that is suited for such
experiments is the polarized Madin Darby canine kidney cell
FIG. 5. Detection of thyroglobulin in the mouse thyroid by immunogold electron microscopy. After wash with phosphate-
buffered saline–bovine serum albumin (PBS-BSA 1%), ultrathin sections (0.1 lm) were incubated overnight with a rabbit
polyclonal anti-thyroglobulin antibody (1/300, DAKO). Sections were then rinsed and incubated for 30 minutes with a 12-nm
colloidal gold affinity pure goat anti-rabbit IgG ( Jackson, 111-205-144, lot no. 71647). Sections were postfixed with 2.5%
glutaraldehyde for 5 minutes and counterstained. They were examined with a Zeiss 109 transmission electron microscope.
(A) Negative control obtained by omission of primary antibody. (B) Thyroglobulin was detected as small gold particles in the
colloid limited by flat epithelial cells. Reproduced with permission from Senou et al. (20).
94 BIANCO ET AL.
line (40). Transfection of these cells is very inefficient and this
may require establishing stably expressing cell lines or viral
transduction with appropriate vectors. Moreover, efflux can be
followed using EYFP H148Q/I152L as an indicator of the in-
tracellular iodide concentration (35–37).
&
RECOMMENDATION 3a
Kinetics of thyroid gland iodide uptake can be studied via
administration of radioactive iodide. Data points can be
obtained in vivo or following en bloc resection of the tra-
chea and thyroid.
Commentary. Thyroid radioactive iodide uptake (RAIU)
and other aspects of iodine kinetics can be studied in rodents
using different iodine isotopes, most commonly
125
I, which
are injected intraperitoneally (2–10 lCi
125
I). The thyroid
gland is subsequently studied at different time points either
with a gamma probe (used, for example, for the identification
of parathyroid tissue in minimally invasive surgery) under
anesthesia (45,46) or dissected postmortem with the trachea en
bloc under a microscope and processed for radiometry for 1
minute in a gamma counter. The results may be expressed as a
function of
125
I in the serum (47) or as percentage of the total
injected dose (46). Thyroid RAIU reaches a maximum at ap-
proximately 4 hours after administration of
125
I and plateaus
at about 12 hours (48). These are approximate time points that
may vary according to the species and strain of the rodent
under investigation. Timing of the
125
I injection can be coor-
dinated with the injection of bovine TSH (bTSH; 10 mU) to
evaluate the TSH-induced thyroidal RAIU. In some settings it
is useful to suppress endogenous TSH by pretreating the an-
imals with T
3
for 4 days prior to radioisotope administration
(48). This will minimize the possibility that endogenous TSH,
which could be different between two groups of animals, is
interfering with the response to bTSH. Notably, a compara-
tive study in rats and mice using recombinant human thyro-
tropin (rhTSH) indicates that it is far more important to
pretreat with T
3
and suppress endogenous TSH in rats than in
mice (49).
&
RECOMMENDATION 3b
Thyroid iodide organification can be quantified via the
perchlorate discharge test.
Commentary. The perchlorate test permits quantification
of the amount of iodide that is normally bound to thyro-
globulin (50). The test is based on the fact that iodide is
transported into thyroid cells by NIS, then released into the
follicular lumen where it is rapidly covalently bound to tyr-
osyl residues of thyroglobulin (organification). Anions such as
perchlorate inhibit NIS, and any intrathyroidal iodide that has
not been incorporated into thyroglobulin is released rapidly
into the bloodstream at the basolateral membrane and cannot
be transported back into thyrocytes. In the standard per-
chlorate test, the thyroidal counts are measured at frequent
intervals after the administration of radioiodine in order to
determine the uptake into the thyroid gland. After doc-
umenting the uptake, perchlorate is administered intrave-
nously or intraperitoneally, and the amount of intrathyroidal
radioiodine is measured at frequent intervals. Under condi-
tions of normal iodide organification, there is no significant
decrease in intrathyroidal counts. In contrast, a loss of 10%
indicates an organification defect, which can be partial (10%–
90%) or complete ( >90%).
In mice, sodium perchlorate (NaClO
4
) is injected intraper-
itoneally 1 hour after injection of
125
I intraperitoneally, and
animals are killed 1 hour later (47). Radioactivity remaining in
the thyroid gland of perchlorate-treated animals is then
compared with the
125
I uptake measured in glands from
control mice that were not exposed to the perchlorate-induced
iodide chase. Protein-bound
125
I (i.e., the total radioactive
thyroid hormones bound to serum transport proteins) is de-
termined in all blood samples after trichloroacetic acid (TCA)
precipitation (47). Others have been able to trace iodide up-
take and discharge in mice directly using gamma probes (45).
Potassium perchlorate (KClO
4
) has been used in rats 6–18
hours following injection of
125
I and shown to reduce the
125
I
thyroid/blood ratio when thyroid peroxidase is inhibited
(51,52).
124
I positron emission tomography/computerized
tomography (PET/CT) has been used rarely to evaluate up-
take and discharge of iodide in rodent thyroids in vivo (53).
&
RECOMMENDATION 3c
Kinetics of thyroidal secretion can be studied in vitro using
en bloc resection of the trachea and thyroid.
Commentary. This strategy is used to evaluate in vitro
TSH-induced thyroidal secretion, minimizing the interference
of other in vivo factors (20). Mice are given an intraperitoneal
injection of about 30 lCi of
125
I and 24 hours later the trachea
and thyroid are removed en bloc and incubated for 3 hours in
Krebs-Ringer bicarbonate medium containing 0.5 g/L bovine
serum albumin (BSA), 8 mM glucose, and 10
- 4
M NaClO
4
to
avoid iodide recirculation. Radiolabeled thyroid hormone
secreted in vitro is extracted with butanol (54). The secretion is
expressed as a percentage of the total radioactivity in the tis-
sue at the beginning of the incubation. There is an approxi-
mately 10-fold induction in thyroidal secretion with the
addition of 5 mU/mL TSH (20).
[A.3] Thyroid imaging
Background. Thyroid imaging in small rodents has fol-
lowed the techniques developed for humans such as scanning
with iodide isotopes, microPET, CT, and high-frequency ul-
trasound (HFUS). However, the minute size of the gland still
poses a significant challenge to obtaining high-quality high-
resolution images, which has been partially overcome by re-
cent new technology.
&
RECOMMENDATION 4a
Thyroid gland functional imaging can be performed using
radioactive iodide isotopes and image acquisition in a
gamma camera or via microPET-CT.
Commentary.
123
I and
131
I can be used together with a
gamma camera for planar imaging as well as single photon
emission computed tomography (SPECT) studies.
131
I has a
long half-life (8 days), but its high energy produces poor
quality images. In contrast, the low energy emitter
123
I is ideal,
producing useful scintigrams with a low absorbed dose; the
main limitations result from its short half-life (13 hours).
Thyroid scintigraphy in anesthetized rats can also be per-
formed 1–24 hours after an intraperitoneal injection of 10 lCi
INVESTIGATING THYROID HORMONE ECONOMY AND ACTION 95
125
I using SPECT (46). Imaging is substantially improved by
placing the animals on a low-iodine diet (LID) for about 3
weeks prior to the studies (46). This enhances the 4 hour
thyroid RAIU from about 3.5% to 27% and makes thyroid
scintigraphy, at all acquisition times, brighter and more de-
tailed (46). SPECT studies in mice using
99m
Tc or
123
I have also
been reported (55,56).
PET studies of the thyroid using
124
I produce good image
quality with a reasonable half-life (4 days). The sensitivity of
PET is higher than that of a scintillation camera, as well as the
contrast and spatial resolution (53). For accurate thyroid im-
aging in rats, the combination of microPET and micro com-
puted axial tomography with
124
I is necessary (Fig. 6) (57).
Anesthetized adult rats or mice are injected via tail vein with
20–540 lCi of Na
124
I and scanned in the microPET for 40
minutes at 24, 48, and 72 hours post injection under anesthesia.
The resulting image data are then normalized to the adminis-
tered activity in terms of the percentage of the injected dose per
gram of tissues (Fig. 6B). Manually drawn 2D regions of in-
terest or 3D volumes of interest can be used to determine the
thyroidal area and volume. For example, the thyroid volume of
an adult 400–500 g rat varies between 35 and 70 lL (57). In
addition to the thyroid gland itself, this approach has also been
used to image metastases of thyroid cancer in mice (58).
&
RECOMMENDATION 4b
Morphological microimaging of the thyroid gland can be
performed by HFUS.
Commentary. HFUS (20–100 MHz) is an imaging meth-
odology that extends the in vivo visualization to microscopic
resolution (of the order of 100 lm; Fig. 7) (59,60). The thyroid
gland of a mouse can be examined using a microimaging sys-
tem that has a single element probe of center frequency and a
dynamic range of 52 dB. HFUS is performed under general
anesthesia (e.g., 1.5%–2% isoflurane vaporized in oxygen) on a
heated stage. Fur is removed from the area of interest (neck and
the high thorax) to obtain a direct contact of the ultrasound gel
FIG. 6. Thyroid imaging using
124
I-iodide in vivo. (A) Biodistribution of
124
I-iodide in thyroid of genetically modified mice in
which thyroid iodide uptake is suppressed by induction of a transgene; 1 week later suppression is relieved and iodide
uptake is normalized. Top panels: representative images of uninduced mice, 1 week on doxycyclin to induce the transgene,
followed by 1 week off doxycyclin. Positron emission tomography (PET) imaging was performed using an R4 microPET
scanner (Concorde Microsystems) with Na
124
I produced on the MSKCC EBCO TR 19-9 (Advanced Cyclotron Systems Inc.)
using 16 MeV protons on a tellurium-124 target. Mice were injected via tail vein with 1.7–2.0 MBq (45–55 lCi) of Na
124
I. Mice
were imaged 24, 48, and 72 hours later under inhalational isoflurane anesthesia (Forane; Baxter Healthcare) at 1 L/min. List-
mode data were acquired for 5 minutes using an energy window of 250–750 keV and a coincidence timing window of 6
nanoseconds, histogrammed into two-dimensional (2D) projected data by Fourier rebinning, and reconstructed by filter back-
projection using a cut-off frequency equal to the Nyquist frequency. The image data were normalized to correct for non-
uniformity of response of the PET, dead-time count losses,
124
I positron branching ratio, and physical decay to the time of
injection, but no attenuation, scatter, or partial-volume averaging correction was applied. (B) Quantification of thyroid
124
I-
iodide uptake in mice treated with the indicated conditions. ***p < 0.001. An empirically determined system calibration factor
(in units of [lCi/mL]/[cps/voxel]) was used to convert reconstructed voxel count rates to activity concentrations. The
resulting image data were then normalized to the administered activity to parameterize images in terms of the percentage of
the injected dose per gram of tissues (%ID/g). Manually drawn 2D regions of interest (ROIs) or three-dimensional (3D)
volumes of interest (VOIs) were used to determined the %ID/g (decay corrected to the time of injection) in various tissues.
Image visualization and analysis were performed using ASIPro VM software (Concorde Microsystems). (C) Representative
gross appearance of thyroid glands at the indicated times. The boundaries of the thyroid are demarcated by dashed lines.
Scale bar: 1 mm. ID/g, injected dose/gram Reproduced with permission from Chakravarty et al. (58).
96 BIANCO ET AL.
to the skin of the animal, minimizing ultrasound attenuation.
Real-time imaging can be performed with a frame rate of 20 Hz
(corresponding to a temporal resolution of 50 milliseconds); the
center of the mouse thyroid is placed about 6 mm from the
transducer’s focal zone. The study, including measurements
and acquisition of accurate, repeatable, and high-quality im-
ages, can be completed in about 30 minutes in the hands of a
well-trained and skilled operator (61).
The volume of each lobe can be calculated using the ovoid
formula (width · depth · length · p/6) (61). The thyroid vol-
ume of an adult C57BL/6 mouse ranges between 2.1 and
4.9 lL. In 6-n-propyl-2-thiouracil (PTU)-treated mice there is
diffuse goiter with volumes that range between 4.1 and 8.8 lL.
Thyroid nodules can be detected via this methodology as
well, with the smallest detectable nodule exhibiting a diam-
eter of 0.46 mm. Features suggestive of malignancy can also be
identified such as hypoechogenicity relative to adjacent nor-
mal tissue, poorly defined margins, internal microcalcifica-
tion, irregular shapes, and extraglandular extension (61). This
should be useful in the phenotypic characterization of mouse
models of thyroid cancer.
[B] Assessing Circulating
and Tissue Thyroid Hormone Levels
Overview. ‘Thyroid status’ of an organism is the sum of
all thyroid hormone signaling events and depends on both
circulating thyroid hormone levels and on local factors
influencing the nuclear concentration of thyroid hormone in
specific tissues. Thyrotoxicosis is the clinical syndrome asso-
ciated with thyroid hormone excess, whereas hypothyroidism
results from thyroid hormone deficiency. At the same time,
individual tissues could be said to have specific thyroid sta-
tus, i.e. hypothyroid or thyrotoxic, relatively independent of
serum thyroid hormone levels; this is because of tissue-
specific deiodinase activities and/or transport mechanisms
(Fig. 8). For example, ischemia and hypoxia cause the brain
and the heart to become acutely hypothyroid in an otherwise
euthyroid animal due to induction of type III deiodinase (D3)
expression (62–65). At the same time, the brown adipose tis-
sue (BAT) exhibits localized increase in thyroid hormone
signaling shortly after rodents are moved to the cold due to
acute induction of type II deiodinase (D2) expression (66).
A common way of assessing thyroid status of an organism,
a.k.a. systemic thyroid status, is by measuring serum levels of
thyroid hormone (T
4
and T
3
) and TSH as well; reverse T
3
can
also be measured, but it is usually reserved for special situa-
tions to confirm abnormalities in thyroid hormone metabo-
lism. Tissue-specific thyroid status can be characterized via
direct measurement of tissue thyroid hormone levels. Typi-
cally, measuring the expression of T
3
-responsive genes (see
Section I) and/or T
3
-responsive biological parameters (see
Section J) is also part of the work up to define thyroid status.
As with the clinical assays developed for patients, a num-
ber of immunoassays for T
4
,T
3
and TSH have been developed
specifically for rodents, which take into consideration differ-
ences in types and capacity of serum iodothyronine binding
proteins and species-specificity of the TSH molecule. In gen-
eral, these assays function well and exhibit sufficient precision
to evaluate thyroid function and systemic thyroid status in
rodents. Under experimental circumstances or specific genetic
defects, serum iodothyronine levels may not reflect thyroid
hormone signaling at the tissue or cellular level. In these cases,
thyroid status can be ascertained by measuring T
3
concen-
trations in specific tissue or cells by adapting the immunoas-
says developed for serum measurements.
[B.1] Serum
Background. Serum thyroid hormone levels may vary
substantially according to sex, age, and strain of the rodent
and should be accounted for in study design. Elevated levels
FIG. 7. High-frequency ultrasonography (HFUS) of the mouse thyroid. (A) Representative image of mouse thyroid using
HFUS and its anatomic correlation with (B) histological transversal images of the subhyoid and tracheal regions. Visible
structures include: 1, tracheal cartilage ring; 2, salivary gland; 3, sternohyoideus and sternothyroideus muscles; 4, sternomas-
toideus muscle; 5, thyroid lobes; 6, common carotid arteries; 7, deep prevertebral muscles scalenus and longus colli; and 8, skin.
A Vevo 770 microimaging system (Visualsonics, Toronto, Ontario, Canada) with a single element probe of center frequency of
40 MHz is used. The transducer has an active face of 3 mm, a lateral resolution of 68.2 lm, axial resolution of 38.5 lm, focal
length of 6 mm, mechanical index 0.14, and a dynamic range 52 dB. A probe with lower frequency and more penetration depth
can also be used (30 MHz center frequency single element with focal depth 12.7 mm, lateral resolution of 115 lm, axial reso-
lution of 55 lm). HFUS is performed under general anesthesia. In this study, mice were anesthetized using 1.5%–2% isoflurane
vaporized in oxygen on a heated stage, with constant monitoring of their body temperature. Area of interest was shaved (neck
and the high thorax) with a depilatory cream to obtain a direct contact of the ultrasound gel to the skin of the animal minimizing
ultrasound attenuation. To provide a coupling medium for the transducer warm gel was used. An outer ring of thick gel
(Aquasonic 100; Parker Laboratories, Orange, NJ) was filled with a thinner gel (echo Gel 100; Eco-Med Pharmaceutical,
Mississauga, Canada) over the region of interest. Reproduced with permission from Mancini et al. (61).
INVESTIGATING THYROID HORMONE ECONOMY AND ACTION 97
(as defined bet ween adjusted normal ranges) usually indi-
cate thyrotoxicosis, while decreased serum levels are indic-
ative of hypothyroidism. Iodine deficiency, alterations in
thyroid hormone metabolism, as well as hypot halamic and
pituitary sensitivity to thyroi d hormone can alter the quan-
titative reciprocal relation between serum T
4
,T
3
,andTSHas
is often the case in mo dels of resistance to TSH or thyroid
hormone.
Immunoassays were developed decades ago and have
served as the cornerstone to measure serum iodothyronines
and T SH. However, the se original assays have been largely
replaced by newer immunoassays (e.g., enzyme-linked im-
munosorbent assay [ELISA], immune radiome tric assay
[IRMA]), all of which are commercially available. Using
commercially available kits to measure serum iodothyr-
onines in rodents is not straightforward because many of
these kits are developed for human serum and m ake use of
an artificial matrix to mimic human binding proteins with
higher affinity and capacity thanthoseofmiceorrats.These
kits utilize ‘displacement agents’ to displace the iodothyr-
onines from human thyroxine binding globulin (TBG; e.g., 8
anilino naphthalen sulfonic acid, di phenyl hyda ntoin, sal-
icylic acid) that are frequently used in excess (for mouse). In
this respect, the y interfere more with T
3
than T
4
,andpar-
ticularly when serum T
3
values are low. This can only
be appropriate ly corrected for by using iodothyronine-
deficient mouse serum as blank and constructing a standard
curve that will calibrate the assay; for e xample, serum f rom
paired box gene 8 (Pax8) knock-out (KO) mice not tre ated
with T
4
or T
3
for at least 2 months (67). Technical limitations
also require the utilization of TSH-deficient mouse serum for
blank and the preparation of a standard curve with mouse
serum TSH, not pituitary TSH, as standard (68). Liquid
chromatography/tandem mass spectrometry is also be-
coming available, although its applicability for rodents is
limited because the required serum volumes are still too
large.
&
RECOMMENDATION 5a
Serum total T
4
and T
3
concentrations can be measured by
radioimmunoassay (RIA), or a host of other immunoas-
says such as ELISA or IRMA, provided that the standard
curves are prepared with rodent serum stripped of thyroid
hormone.
Commentary. Typical standard curves are prepared over
the range 2.5–240 ng/mL for T
4
and 0.1–6 ng/mL for T
3
. These
assays can be developed in house by modification of kits for
human use obtained from multiple commercial sources.
Homemade RIAs have a greater sensitivity with measure-
ments over the range of 0.05–3 ng/mL (67). Clinical assays
developed for patients can be used as long as the rodent
standard curve is parallel to the standard curve provided in
the kit and an appropriate correction factor applied; the ro-
dent standard curve should be used to calculate the results
(67). Commercially available kits designed for measurement
of mouse serum T
3
and T
4
in 10 lL samples have been de-
veloped and used with acceptable results (69).
&
RECOMMENDATION 5b
Assays for measuring circulating T
4
and T
3
are best per-
formed using serum rather than plasma, since fibrin for-
mation affects pipetting, and additives such as heparin may
directly interfere with free hormone determination.
Commentary. Frequent blood samples can be obtained
during the course of an experiment if limited to approxima-
tely 10% of the total volume every 2–4 weeks and 1% every 24
hours. Serum can be stored at -20C for long time periods.
The use of anesthesia may have variable effects on thyroid
hormone levels, and each investigator should evaluate po-
tential effects in their system with the anesthetic they are
using. Serum T
3
and T
4
exhibit minimal circadian variations
along day–night cycles; these could be taken into account
depending on the timing of sample collection. Serum samples
FIG. 8. Supply and metabolism of thyroid hormones affect negatively and positively T
3
-regulated genes in the brain. To
construct this figure, the authors used individual reverse transcriptase quantitative polymerase chain reaction (RT-qPCR)
data from T
3
-regulated genes to calculate the fold change relative to the wild-type (WT) values, and plotted the Log
2
FC (fold
change) to make the results quantitatively comparable. The data were represented in a box-and-whiskers (5%–95%) plot.
Statistical significance between each group and the WT was calculated by one-way ANOVA. For the positive genes,
F
5,537
= 272, p < 0.0001. For the negative genes, F
5,400
= 145, p < 0.0001. *p < 0.05; **p < 0.01; ***p < 0.001. Reproduced with per-
mission from Hernandez et al. (492).
98 BIANCO ET AL.
with milky aspect from lactating dams or from their pups can
give erroneous results due to their high lipid content. In these
cases extraction of the serum and removal of the lipids using
chloroform is advisable (67).
&
RECOMMENDATION 5c
Determinations of free iodothyronine indexes (FT
4
I and
FT
3
I) in the serum can be achieved by measurement of the
total serum hormone concentration and the serum io-
dothyronine binding capacity using one of the resin or
charcoal methods.
Commentary. The existence of proteins in the serum
that reve rsibly bind thyroid hormone establishes two pools
of circulating T
4
and T
3
(i.e., protein-bound and free). The
major circulating high affinity thyroid hormone binding
proteins differ in rodents and humans, with transthyretin
being the major protein in the rat and TBG in humans. It is
the free thyroid hormone in the plasma that is in equilib-
rium with tissues and affects thyroid hormone signaling.
Measurement of free hormone by methods other than
equilibrium dialysis can give erroneous results, though
microfiltration of the samples has been used with reliable
results (70,71). Equilibrium dialysis of serum with labeled
iodothyronine tracer in dialysis bags has been used to
measurefreeT
4
and T
3
in the rat. The method is not used for
mice, owing to the requirement of more than 1 mL of serum
for measurement in triplicate because leaks often occur.
Using diluted serum and applying correction is not advis-
able. Only 100 lL of serum is required when using micro-
filtration of the samples (70).
Alternatively, an estimate of the FT
4
IorFT
3
I can be ob-
tained using a relatively small volume of serum by using
the resin or charcoal methods (72). Serum is diluted into
phosphate-buffered saline (PBS; pH 7.4) containing [
125
I]T
3
or
[
125
I]T
4
. Samples are allowed to equilibrate and subsequently
mixed with 0.0125% activated charcoal solution. Charcoal
pellets are obtained and then counted in a c-counter. Condi-
tions should be optimized such that approximately 20%–30%
of the tracer is bound to charcoal in sera from euthyroid
control animals. An estimate of the free T
4
or T
3
(FT
4
IorFT
3
I)
can be calculated by multiplying the total T
4
or T
3
serum
concentration by the T
4
or T
3
charcoal uptake.
&
RECOMMENDATION 5d
Isotope dilution tandem mass spectrometric can be used to
measure T
4
and T
3
in biological samples.
Commentary. Immunoassays for thyroid hormone mea-
surement can suffer from poor specificity. As an alternative,
simultaneous measurement of T
4
and T
3
can be achieved by
using isotope dilution tandem mass spectrometry within a
single run (67,73). The method requires 100 lL of serum and
involves addition of internal standard, precipitation of pro-
teins with methanol and injection onto a C-18 column. T
4
and
T
3
are subsequently eluted using a methanol gradient. This
method is accurate, specific, and precise (coefficient of varia-
tion of 3.5%–9.0%). A concern is the sample volume needed
for free hormone determination, which is still relatively large
for applications involving mice, except in terminal bleeding.
Similar methodology applying liquid chromatography-
tandem mass spectrometry has been developed for measure-
ment of iodothyronamines, a decarboxylated iodothyronine
present in a number of biological fluids (74).
&
RECOMMENDATION 6a
Rat and mouse serum TSH can be measured using com-
mercially available rat TSH assay kits. Alternatively,
species-specific RIAs can be performed using reagents from
the National Hormone and Peptide Program, National In-
stitute of Diabetes and Digestive and Kidney Diseases
(Bethesda, MD).
Commentary. In general, RIAs for TSH are more sensitive
than IRMAs. Commercial assays do not provide species spe-
cific cross reference and, therefore standard curves are rarely
parallel to values obtained with actual sample dilution.
However, commercial reagents can be adapted for specific
and accurate measurements of TSH as outlined below.
TSH standard curves should be constructed using species-
specific circulating (serum, not pituitary) TSH standard,
diluted in TSH-deficient serum obtained from the same
species. Serum TSH standard is obtained from animals ren-
dered hypoth yroid (see Section E.1). The content of TSH is
calibrated against a b TSH standar d in a bioas say. TSH-
deficient serum is prepared by making roden ts thyrotoxic
(treatment with 20 lg levothyroxine [L-T
4
]/100 g body
weight [BW]/day for 1 week). Sample nonparallelism with
standard curves is due to species differences and to cross-
reactivity with free TSH subunits and other pituitary gly-
coproteins in pituitary extracts. The use of lactoperoxidase to
label TSH with
125
I improves the stability of the labeled TSH
and the sensitivity of the TSH assay up to thyrotoxic ranges
(68). Measurement of TSH concentration in pituitary gland
extracts can be done, however, using the same assay at a
dilution of 1:500 to 1:2000 in assay buffer. The standard
curve can also be built using buffer, instead of TSH-deficient
rodent serum, as the diluent. Running the RIAs in d isequi-
librium (addition of the isotope t racer for a shorter time after
incubation of the TSH antibody with the samples) improves
the sensitivity of the assay. If no reliable rat/mouse serum
TSH measurement is available, the levels of TSHb mRNA in
the pituitary gland can be used as an indication of TSH
production (75).
&
RECOMMENDATION 6b
TSH biological activity can be studied by standardized
in vitro assays as well as in vivo assays.
Commentary. TSH biological activity is modulated by a
number of factors including its structure, glycosylation or
carbohydrate branching, as well as by the TSH receptor. The
biological activity of the TSH molecule can be determined by
an in vitro bioassay using Chinese hamster ovary cells stably
expressing the TSH receptor (68,76). The subclone cl 213 of
JP2626 is particularly sensitive to low levels of TSH. About
50,000 cells are seeded in individual test tubes and incubated
with 20 lL of serum, followed by cAMP extraction with 0.1 M
HCl and measurement by RIA (77). Blanks are processed as
already described with TSH-depleted serum obtained from
T
4
-treated mice. cAMP production is a function of how much
endogenous TSH was contained in the plasma sample (68).
Dividing the cAMP generated in vitro by the TSH values in the
plasma sample provides an index of TSH biological activity.
INVESTIGATING THYROID HORMONE ECONOMY AND ACTION 99
Of course, this can vary according to mutations in the TSH
molecule or degree and type of glycosylation or carbohydrate
branching. However, such changes may not always show
biologic differences using in vitro tests. Alteration in the pro-
tein glycosylation or the tertiary structure (carbohydrate
branching) of the sugar residue, alters the half-life of TSH
in vivo and affects its bioactivity. However, this cannot be
always demonstrated by in vitro bioassay. In some instances it
can be shown by isoelectric focusing, if a sufficient amount of
TSH can be concentrated and developed by Western blotting
or by Concanavalin-A chromatography (78). Another method
is to affinity purify the TSH being tested, inject it intrave-
nously in TSH-suppressed mice or rats (treated with high
dose of T
3
), and follow its half-life by RIA, or follow the bio-
logical activity of TSH by measuring T
4
secretion in serum.
Decreased TSH bioactivity can also be caused by defects in the
TSH receptor or reduced number of TSH receptors expressed
in the follicular cell surface (e.g., heterozygous TTF1 KO
mice). This can be confirmed by showing intact response in an
in vitro bioassay along with alteration in the response of the
animal to injected authentic TSH (79).
&
RECOMMENDATION 7
Thyrotropin releasing hormone (TRH)-induced TSH se-
cretion testing can be used to assess the capacity of the
pituitary gland to secrete TSH. TSH-induced thyroidal se-
cretion testing can be used to assess the capacity of the
thyroid gland to produce and secrete thyroid hormone.
Commentary. TheTRH-TSHaxiscanbeinterrogatedat
either the pituitary or thyroid glandular level via specific dy-
namic tests. The TRH stimulation test is performed with an
intravenous or intraperitoneal injection of TRH (5.0 lg/kg BW).
Blood is collected 30 minutes later for measurement of serum
TSH and 2 hours later for measurement of serum T
3
,which
indicates the thyroidal responsiveness to TRH-induced TSH.
The expected increase in serum TSH is about threefold, whereas
an elevation of approximately 50% in serum T
3
is expected (80).
The TSH stimulation test is performed with an intravenous
or intraperitoneal injection of bTSH (2–250 mU/100 g BW).
Two hours later, blood is collected for measurement of serum
T
3
, with an expected elevation of approximately 40% com-
pared to baseline levels (79–81). An alternative approach is to
pretreat mice for 4 days with T
3
(1 lg/d) in order to suppress
endogenous TSH and then administer bTSH (2, 10, or 30 mU)
on the morning of the fifth day. In this case, the thyroidal
response is evaluated based on the TSH-induced elevation in
serum T
4
3 hours later, which is about 1 lg/dL (81) or three-
fold over baseline (82). A similar approach can be used in rats,
and the TSH-induced T
4
response varies quite substantially
according to the rat strain. Still in rats, the TSH-induced T
4
response plateaus at a bTSH dose of about 100 mU, with an
increase of about 2 lg/dL above baseline (79). Of note, a
comparative assessment of the thyroid responsiveness to
rhTSH in rats and mice indicates poor or no response in rats
that were not pretreated with T
3
(49).
[B.2] Tissue
Background. While the plasma constitutes the largest
extrathyroidal pool of T
4
, approximately two thirds of all T
3
is
found in the intracellular space and initiates thyroid hormone
action by binding to nuclear thyroid hormone receptors (TRs).
The intensity of the signaling depends on the number of oc-
cupied TRs in any given T
3
-responsive tissue. Because the
extracellular and intracellular compartments are in commu-
nication and thyroid hormone molecules transit in and out of
the cells via the different membrane transporters, in most
tissues measuring the serum concentration of thyroid hor-
mone provides an estimate of the intracellular T
3
concentra-
tion. However, a disruption of the transport system might
prevent free access of T
3
to the intracellular compartment. In
addition, intracellular metabolism of thyroid hormone, both
activation and inactivation, might affect thyroid hormone
signaling in a way that cannot be predicted from sampling the
plasma compartment. Thus, serum levels of T
3
do not neces-
sarily reflect the amount of T
3
in all tissues or the intensity of
thyroid hormone signaling. Direct measurement of tissue T
3
content provides this additional information.
&
RECOMMENDATION 8
Tissue content of T
3
and T
4
can be measured by immuno-
assays after tissue extraction.
Commentary. Removing blood from tissues by perfusion
is important particularly for highly vascular tissues. After
collecting a blood sample, mice are perfused with heparin
containing PBS through a needle placed in the left ventricle
(LV) of heart followed by cutting open the vena cava. Tissues
are then collected, immediately frozen on dry ice, and stored
at -80C. Iodothyronines are extracted from tissues using
methanol–chloroform (1:2). The amount of tissue to be ex-
tracted depends on thyroid hormone status of the animal and
the hormone abundance in a specific tissue. As an example,
50 mg of brain and 15 mg of liver of an euthyroid mouse will
generally yield satisfactory results, but the amount should be
increased in samples from hypothyroid mice or rats. Radio-
active T
3
or T
4
should be added to each sample to determine
efficiency of extraction; a mix of [
125
I]T
3
and [
131
I]T
4
can be
used when both hormones are to be studied. Depending on
the extraction procedure, chloroform should be removed be-
cause it contains lipids and other substances that interfere in
the RIAs. This involves back-extraction in calcium chloride,
concentration of the extracts, and evaporation (83). Once ex-
traction is completed, the dried extract is dissolved, preferably
in buffer or charcoal-stripped rodent serum, and T
3
content
measured by the specific immunoassay, following the given
recommendations. A highly sensitive RIA is decisive to obtain
reliable results in small samples or in samples from hypo-
thyroid animals. For determination of tissue T
4
content,
commercial assays are not sensitive enough and a highly
sensitive T
4
RIA in buffer should be used (83). All assays must
include appropriate blank/control tubes, containing all re-
agents except for the tissue sample, to be used to check the
assay background. Validation of the assay also includes
demonstrating parallelism between a tissue curve (multiple
points with progressively greater amounts of tissue extract)
and the standard curve over the range of interest.
[B.3] Sources of tissue T
3
and TR saturation
Background. T
3
present in extrathyroidal tissues may be
derived from two distinct sources: plasma T
3
and T
3
locally
generated from T
4
(84,85). The latter mechanism is typically
100 BIANCO ET AL.
found in tissues that express D2 such as brain, pituitary, and
BAT. Estimates suggest that at least half of the T
3
present in
D2-expressing tissues is produced locally from deiodination
of T
4
(86–89). More recently D2 expression has been found in a
large number of tissues and cells (90–98), illustrating the im-
portance of defining its contribution to tissue-specific thyroid
hormone signaling. The determination of the sources of in-
tracellular T
3
is feasible because plasma T
3
equilibrates rap-
idly with most tissues (but not all). At the equilibrium time
point (Tm) one can use the plasma T
3
concentration and the
nuclear/plasma ratio of tracer T
3
to estimate the amount of
nuclear T
3
that is derived from plasma. A similar strategy can
then be applied for T
4
, returning the nuclear T
3
that is derived
from local conversion of T
4
to T
3
.
&
RECOMMENDATION 9a
The contribution of plasma T
3
to tissue T
3
can be quantified
by tissue-labeling techniques involving either single intra-
venous injections or pump-driven chronic infusion of
radiolabeled tracer T
3
.
Commentary. Studies have been standardized in rats but
could in theory be applied in mice as well, provided that
limitations due to body size and anesthesia are overcome.
Tissues can be studied as a whole or fractionated to isolate the
TR-containing nuclear fraction (66,86–89). After the adminis-
tration of [
125
I]T
3
, Tm is defined as the time at which the
amount of tracer [
125
I]T
3
entering the tissue or nuclear com-
partment equals the amount of [
125
I]T
3
exiting the same
compartment. Tm is reached within hours of the intravenous
injection or within days of the pump start. At the Tm, the
[
125
I]T
3
plasma/tissue ratio and the plasma concentration of
T
3
are used to calculate the tissue T
3
concentration. Similar
calculations are used in case radiolabeled tracers are infused
via pumps (99,100). These methods have been standardized
with
125
I-T
3
separation by descending paper chromatography.
There is good agreement that high performance liquid chro-
matography (HPLC) and ultra performance liquid chroma-
tography (UPLC) are excellent methods for separating labeled
iodothyronines and in theory could be used as well.
&
RECOMMENDATION 9b
TR maximum binding capacity in a tissue can be estimated
via saturation analysis with T
3
and data reduction using the
Scatchard method.
Commentary. The combined administration of tracer
[
125
I]T
3
with increasing amounts of cold T
3
progressively
saturates the high affinity T
3
binding sites (TR) (86–89,101). In
this case, nuclei are isolated and processed for [
125
I]T
3
content.
The plasma T
3
concentration and the plasma/nuclear ratio at
the Tm are then obtained for each dose of cold T
3
that was
injected. Results are expressed per milligram of DNA, and the
Scatchard analysis of the data allows for the calculation of the
TR maximum binding capacity and relative affinity in any
given tissue. The plasma T
3
versus nuclear T
3
curve makes it
possible to calculate the TR saturation at any given level of T
3
,
including physiological plasma levels.
&
RECOMMENDATION 9c
Dual-labeling techniques using [
131
I]T
3
and [
125
I]T
4
can be
used to determine the relative contributions of plasma T
3
versus locally produced T
3
via T
4
deiodination to tissue T
3
concentration.
Commentary. The administration of [
125
I]T
4
and subse-
quent measurement of plasma and tissue [
125
I]T
3
allows for
the quantification of locally produced T
3
in tissues as a whole
or TR-containing nuclear fraction (86–89). Even if relatively
large activities of [
125
I]T
4
are used, the amounts of [
125
I]T
3
produced at the Tm are minimal in euthyroid animals. Thus,
both plasma and tissue (nuclear) [
125
I]T
3
should be concen-
trated using an anti-T
3
affinity column before separation
by chromatography. At the Tm, the plasma/tissue ratio of
[
125
I]T
3
/[
125
I]T
4
and the serum T
4
concentration are used to
calculate the locally produced T
3
. Values of [
125
I]T
3
are mul-
tiplied by 2 given that there is only one
125
I in the phenolic
(outer) ring of [
125
I]T
4
and deiodination occurs randomly be-
tween the 3¢ and 5¢ positions. Appropriate corrections should
be used when the tracer contains radioactive iodine in both
phenolic (outer) and tyrosil (inner) rings. To account (and
discount) for the contribution of plasma [
125
I]T
3
(exiting from
tissues) to tissue [
125
I]T
3
, administration of [
125
I]T
4
is coordi-
nated with the administration of [
131
I]T
3
. Because the Tm(T
4
)
and Tm(T
3
) are different, the administration of the two tracers
should be timed so that both Tms coincide at the time animals
are killed.
[C] Assessing Thyroid Hormone Transport into Cells
Overview. Based on the lipophilic structure of thyroid
hormones, it was long thought that thyroid hormone enters
the cell through passive diffusion. However, it has become
clear that thyroid hormones are transported across the plasma
membrane via carrier-mediated transport, providing the cell
with an important tool to regulate intracellular thyroid hor-
mone availability. Carrier-mediated transport of thyroid
hormones is facilitated by specific substrate–transporter in-
teractions and predominantly driven down concentration
gradients and, for some transporters, through the co-transport
of other molecules. Solute carriers known to transport thyroid
hormones include monocarboxylate transporters (MCTs),
Na
+
/taurocholate co-transporting polypeptide, organic anion
transporters (OATs), amino acid transporters (e.g., L-type
amino acid transporters), and organic anion transporting
polypeptides (OATPs) (102–106). The importance of thyroid
hormone transporters is illustrated by the fact that mutations
in human MCT8 cause psychomotor retardation and altered
iodothyronine levels (107,108). With the exception of
OATP1C1 (T
4
and 3,3¢,5¢-triiodothyronine [rT
3
] transport
only), most thyroid hormone transporters transport both T
4
and T
3
. It is important to realize that members of the MCT
family, such as MCT8 and MCT10, facilitate not only the
cellular uptake, but also the efflux of iodothyronines (109).
The physiological role of the transporters is not only depen-
dent on their relative affinities for the thyroid hormones but
also depend upon tissue- and cell-specific expression patterns.
A confounding factor in teasing out the functional role of
specific transporters is the possible expression of different
transporters on the surface of an individual cell type (e.g.,
hepatocytes and neurons). In these instances the relative
abundance of a specific family member will likely determine
the hierarchy of transport functions for a specific cell type.
Finally, several thyroid hormone transporters do not
INVESTIGATING THYROID HORMONE ECONOMY AND ACTION 101
transport thyroid hormones exclusively. For example, the
OATPs transport a wide variety of both endobiotics and xe-
nobiotics and possess multiple substrate binding sites
(110,111). This complexity in expression and function reveals
some of the challenges associated with studying thyroid
hormone transport both in vivo and in vitro that must be
considered by the experimentalist.
[C.1] Thyroid hormone transport in vitro
Background. In vitro studies are generally considered for
the biochemical characterization of individual transporters,
study of the effects of specific mutations on transporter
function, and elucidation of the role of specific transporters
expressed in individual cell types. A primary consideration
is assurance that the experiment is properly designed and
controlled to test the function of specific transporters versus
the collective action of multiple transporters expressed in
the same cell. Kinetic studies require consideration of possible
confounding factors such as assurance of plasma membrane
localization of the expressed transporter, bidirectional sub-
strate transport, choice of experimental cell type, and typical
versus atypical kinetics.
&
RECOMMENDATION 10a
Iodothyronine transport into cells can be studied in cells
transiently expressing wild-type or mutant thyroid hor-
mone transporters.
Commentary. Assessing functionality of cloned wild-
type or mutant thyroid hormone transporters is readily
achieved by conducting uptake experiments (110,112). Proper
controls include comparison to cells transfected with empty
vector and inhibition of labeled thyroid hormone uptake by
co-incubating with excess cold hormone. The latter control is
critical to ensure that the observed cell associated thyroid
hormone uptake is a saturable process and is not simply as-
sociated with nonspecific binding of labeled hormone to the
cell. For general transport assays, uptake commences with the
addition of labeled substrate to the cells and terminates at
specific time points with rapid washes with cold transport
buffer. These wash steps are essential, since iodothyronines
tend to adhere to the cell walls. Attention should be given to
the selection of transport buffer (e.g., Krebs-Henseleit buffer,
Dulbecco’s modified Eagle’s medium/F12, or regular PBS
with/or without BSA). It is important to realize that different
transport buffers can contain large amounts of amino acids,
which may also be transported by certain transporters and
thereby influence the results of uptake experiments. BSA can
be added to the medium to keep iodothyronines in solution
and prevent the adsorption to plastics. However, by de-
creasing the free iodothyronine concentration it may also limit
substrate availability.
Transport buffers preferably contain no serum at all, since
even ion-exchange resin-stripped serum may still contain low
levels of thyroid hormones interfering with the uptake assay.
As a consequence, kinetic measurements (e.g., transport
constants, maximal velocities) cannot be determined. Uptake
is calculated from the proportion of radioactivity associated
with the cell lysate compared with the total radioactivity as-
sociated with the isotopic transport buffer. Background ra-
dioactivity in cells transfected with empty vector is subtracted
from all samples. Uptake is calculated from the proportion of
radioactivity associated with the cell lysate compared with
total radioactivity associated with the isotopic transport buf-
fer and expressed in units of picomoles per minute (113,114).
Different results may be obtained in different cell types,
depending on the endogenous expression levels of thyroid
hormone transporters and/or other factors necessary for
thyroid hormone transport. When studying the function of
mutant transporters, it may be useful to compare results in
cells with high versus low endogenous expression levels.
Thyroid hormone transporters exhibit bidirectional trans-
port. For some transporters such as the OATPs, an anti-ported
substrate is thought to be required for thyroid hormone
transport across the plasma membrane and should be con-
sidered when choosing a transport buffer. Accumulation of
transported thyroid hormones in the cell is necessary for as-
sessing transport activity and can present a problem for some
bidirectional transporters because the substrate may rapidly
efflux thyroid hormones from the cell. Co-transfection with
intracellular thyroid hormone binding proteins such as mu-
crystallin provides a method for ensuring accumulation of
transported hormone (109,115,116). One caveat, however, is
that use of such methods precludes subsequent kinetic studies
because hormone transport will likely not ever reach steady-
state.
To study the consequences of overexpression of a specific
thyroid hormone transporter on intracellular availability of
thyroid hormone, cells can be co-transfected with an io-
dothyronine deiodinase and subsequently analyzed for me-
tabolism (114). Alternatively, Xenopus oocytes can be used for
transport studies (118).
&
RECOMMENDATION 10b
Determining the substrate specificity and kinetic charac-
teristics for thyroid hormone transporters requires com-
mitment to testing various substrates at a wide range of
concentrations.
Commentary. The substrate specificity of (putative) thy-
roid hormone transporters can be investigated by incubation of
cells with different putative radioactive ligands, including the
different iodothyronines T
4
,T
3
,rT
3
, and 3,3¢-diiodothyronine
(T
2
), as well as thyroid hormone analogues such as 3,5,3¢-
triiodothyroacetic acid (tiratricol, also known as TRIAC) and/
or various amino acids such as Leu, Phe, Tyr, and Trp. Speci-
ficity of TH transport can be measured by determining the
uptake of radiolabeled iodothyronines in the presence of
putative competitors, including unlabeled iodothyronine de-
rivatives such as D- and L-iodothyronines, tiratricol, tetra-
iodothyroacetic acid (Tetrac), and amino acids.
Characterization of kinetic parameters of cloned wild-type
and mutant thyroid hormone transporters requires first de-
termining the time course of substrate influx into the trans-
fected cells. From these data a time point is identified when
substrate uptake is close to maximal but transport has not yet
reached steady-state. All further kinetic studies are then
conducted by measuring substrate accumulation in trans-
fected and control cells at this identified time point.
For analysis of transporter mutants it would be expected
that mutants may possess altered kinetic activities. Therefore,
for the time course study the investigator should assess up-
take at multiple time points over the course of 120 minutes.
102 BIANCO ET AL.
Mutants can be subjected to Michaelis–Menten kinetic anal-
ysis to quantify potential changes in thyroid hormone trans-
port kinetics.
For the subsequent kinetic studies, a time point should be
chosen prior to steady-state when maximal differences be-
tween expressed transporter versus empty vector uptake are
observed. At this time point, one can measure the uptake
using multiple different substrate concentrations spanning
the dynamic range of detection for each substrate. Since these
large experiments are difficult to perform, many investigators
choose to limit the number of data points collected, accepting
a less accurate estimation of transport kinetics. Notably, sa-
turability and apparent Km values may be influenced by
binding of iodothyronines to proteins such as BSA in the
medium.
Assessment of OATP function is even more difficult as
these transporters possess multiple substrate binding sites
and thus exhibit atypical transport kinetics (110,119). To
properly assess OATP transport kinetics gather kinetic data as
described above and use additional methods to analyze the
data (110,120–122). These transport kinetics cannot be deter-
mined for transporters with high efflux capacity, such as
MCT8 and MCT10, which may require co-transfection with
intracellular binding proteins (see previous discussion).
&
RECOMMENDATION 10c
Assessing thyroid hormone transport in dissociated pri-
mary cell cultures in vitro is only of limited value if mixed
cell cultures are used.
Commentary. A caution when working with primary
cells is potential rapid down-regulation of transporter ex-
pression (123). Tracking transporter expression by Western
blot or immunocytochemistry over time in culture is therefore
critical. In addition, the presence of a mixed cell population
mandates identification of cell types expressing specific
transporters (124). However, lack of antibody specificity may
pose significant problems for tracking the expression of spe-
cific members of large transporter families with multiple
closely related members.
&
RECOMMENDATION 11
Confirmation of proper plasma membrane localization is
necessary for studies utilizing cells transiently expressing
thyroid hormone transporters.
Commentary. Ensuring that the assessed transporters are
localized to the plasma membrane is of key importance for
in vitro studies. When appropriate antibodies are available
Western analysis should first be performed to ensure protein
expression followed by fluorescent immunocytochemistry to
assess subcellular protein localization. When studying trans-
fected cells, it is important to study cells with low levels of
endogenous expression of the transporter of interest. For ex-
ample, JEG3 cells are an excellent model to study plasma
membrane localization of mutated MCT8 or MCT10 because
of their low levels of endogenous expression of MCT8 and/or
MCT10. For biochemical analysis, epitope-tagged cloned
transporters provide a ready antigenic target for both Western
blot and immunocytochemical analysis. Subcellular localiza-
tion is preferably determined by confocal microscopy because
typical wide-field fluorescent microscopy will not readily
distinguish between cell surface and intracellular labeling.
Appropriately expressed protein is primarily localized in the
plasma membrane with limited intracellular staining. This
analysis is crucial for proper evaluation of genetic mutants as
a lack of transport observed in in vitro assays could be either
due to a lack of transport function or an inability to properly
traffic the protein to the plasma membrane. In both scenarios
the transport assay would demonstrate a loss of transporter
function, although the biochemical explanation for this result
would be markedly different. Note that these studies may be
difficult to perform in transiently transfected cells, due to a
relatively low transfection efficacy. For such studies the use of
stably transfected cells may be more useful.
[C.2] Thyroid hormone transport in vivo
Background. In vivo study of thyroid hormone transport
is considered for elucidating the physiological and patho-
physiological roles of specific transporters and for determin-
ing the kinetics of tissue influx and efflux of thyroid
hormones. Kinetic studies require consideration of possible
confounding factors such as first pass metabolism of injected
tracer hormone, the presence of multiple thyroid hormone
transporter family members in a single tissue or even cell type,
and developmental changes in transporter expression pat-
terns. Consequences of defective thyroid hormone transport
on thyroid hormone signaling can be analyzed using genome
wide analysis (Section I.3) on patient material such as fibro-
blasts (125), whereas consequences for different systemic
biological parameters can be analyzed using different knock-
out mouse models [see Sections I.5 and J; see Heuer and
Visser (126) for a review].
&
RECOMMENDATION 12a
For in vivo studies, net thyroid hormone transport typically
represents the summation of a number of different trans-
porter activities. Assessing the contribution of a given
transporter to net transport requires examination of all
transporters expressed in the tissue.
Commentary. In vivo measures of thyroid hormone
transport are confounded by the expression of multiple thy-
roid hormone transporters in most tissues. A key consider-
ation is to determine the transporter expression profiles and
levels in specific cell types in targeted tissues. Since the affinity
for T
4
and T
3
is similar across most transporters, the contri-
bution of an individual transporter generally follows its
expression levels. Experimental design should take the pos-
sibility of both hormone influx and efflux into consideration
[see Heuer and Visser (126) for a review]. Consequences of
defective thyroid hormone transport for thyroid hormone
signaling can be analyzed using genome wide analysis (see
Section I.3) of patients’ material such as fibroblasts (125) or
different tissues of knock-out mice (127).
&
RECOMMENDATION 12b
Metabolism of assessed thyroid hormones must be con-
sidered when measuring transport in vivo.
Commentary. Thyroid hormones are subjected to me-
tabolism in vivo including deiodination and conjugative me-
tabolism. Importantly, such metabolism impacts the transport
INVESTIGATING THYROID HORMONE ECONOMY AND ACTION 103
fate of the hormones and when the thyroid hormones me-
tabolized are radiolabeled tracer substrates, interpreting ob-
tained results is difficult. Control over metabolism can be
exerted if single pass strategies are used such as direct can-
nulation of vessels leading to target tissues followed by con-
tinuous infusion of defined transport buffer and substrate.
Such strategies limit metabolism of the substrate and loss to
peripheral tissues and allow direct measurement of thyroid
hormone transport in the tissue assessed. As an example,
measurement of thyroid hormone transport across the blood–
brain barrier would be best achieved using the in situ brain
perfusion method (128). This methodology allows accurate
kinetic measurements such as transport rates and transport
affinities and can be conducted in neonatal to adult animals.
Comparison of transport kinetics in wild-type and specific
transporter null animals would allow determination of the
specific contributions of individual transporters in vivo.
[D] Assessing Thyroid Hormone Deiodination
Overview. The major circulating iodothyronine is T
4
.
However, T
4
can be converted in the tissues to T
3
, the prin-
cipal thyroid hormone that binds to the nuclear receptors and
initiates thyroid hormone action (129). In extrathyroidal tis-
sues, the concentration of T
3
in the intracellular and nuclear
compartments is determined in part by the rates of T
4
to T
3
conversion and T
3
and T
4
degradation in the cell (84,130–132).
The formation and degradation of T
3
in tissues are depen-
dent primarily on the activities of three integral-membrane
thioredoxin-fold–containing selenoenzymes of approxima-
tely 60 kDa (dimer) that catalyze the selective removal of
iodine from iodothyronines (132–135). The type I deiodinase
(D1) and D2 enzymes are activating enzymes that catalyze
the outer-ring or 5¢-deiodination (5¢D) of T
4
to T
3
.D3is
an inactivating enzyme that catalyzes the inner-ring or 5-
deiodination (5D) of both T
4
and T
3
to their relatively inactive
derivatives, rT
3
and T
2
, respectively. The D1 can also inacti-
vate the thyroid hormones by catalyzing the inner ring deio-
dination of sulfated iodothyronine conjugates (130–132).
Levels of activity of D1, D2, and D3 can be studied in cultured
cell preparations (136,137), tissue slices (138,139), and tissue
homogenates (140,141).
[D.1] Identification, expression, and quantification
of deiodinases
Background. The presence of the deiodinases can be
identified in a given tissue and their levels of activity quan-
tified using well-established assays for enzyme activity and
kinetic properties. In the cases of D1 and D3, tissue content of
deiodinase protein can be assessed by immunohistochemical
techniques using available antisera/antibodies (142). Levels
of expression of all three deiodinase genes can be determined
using samples of RNA prepared from the tissue of interest by
standard reverse transcriptase quantitative polymerase chain
reaction (RT-qPCR) techniques (127).
Assay of levels of deiodinase activity and/or the enzyme
reaction kinetics in broken cell/tissue preparations requires the
presence of compounds containing free sulfhydryl groups such
as dithiothreitol (DTT). Assay of D1, D2, and D3 activity levels
are generally carried out using radioactive iodothyronines as
substrates followed by quantitation of the radioactive products
generated per unit time. Alternatively, in the case of the D1 and
the D2, activity levels can be determined using nonradioactive
T
4
and quantifying the T
3
generated by RIA.
&
RECOMMENDATION 13a
Deiodinase activities and their kinetic profiles and intrinsic
properties (e.g., Vmax, Km, activity half-life, sensitivity to
cofactors and/or inhibitors) can be determined in cell or
tissue preparations.
Commentary. There are multiple protocols for deiodinase
assays, and they are all acceptable provided that some re-
quirements are followed. In general, release of tracer radio-
active iodide or generation of a specific deiodinase product
are the experimental endpoints measured. It is always best to
use radiolabeled iodothyronines with the highest available
specific activity. Even when they are stored in the dark at 4C,
there is spontaneous iodothyronine deiodination. Thus, in
order to minimize background levels, tracers should be pu-
rified in Sephadex LH50 at least 24 hours before the assay.
All assays should contain background controls in which there
is no deiodinase-mediated substrate deiodination. Ideally,
duplicate samples are run in parallel in tubes containing a
large excess of substrate that outcompetes the radiolabeled
iodothyronine.
Assay for D1 activity. The preferred substrate for D1 is rT
3
;
taking the Vmax/Km ratio as a measure of efficiency, rT
3
performs as much as 700 times better than T
4
as a substrate for
D1. In addition, unlike T
4
,rT
3
it is a relatively poor substrate for
D3 and thus is the better substrate with which to assay D1 in
tissues that also contain D3 activity. At the same time, D1 ac-
tivity can also be determined using T
4
as substrate. The reaction
buffer varies somewhat among laboratories but most are car-
ried out in a PBS or 2-amino-2-hydroxymethyl-1,3-propanediol
hydrochloride (Tris-HCl) buffer containing ethylenediamine-
tetraacetic acid (EDTA; 1–2 mM), sucrose (0.25–0.3 M), and
DTT (10–50 mM). The substrate concentration is 0.1–2 lM. The
amount of tissue present must be adjusted to ensure a relatively
low percent deiodination (<30%) in order to avoid a significant
alteration of the enzyme kinetics. With broken cell prepara-
tions, the incubation time should be 1 hour or less. When
deiodinase activity is studied in tissue slices or whole cells, the
tissue or cells should be homogenized on ice in their incubation
media prior to analysis of the products, since iodothyronines
tend to adhere to cell walls.
The iodinated products can be quantified in various ways.
Arguably the simplest method is to pass the reaction mixture
through a column of Biorad AG 50W-X8(H+) resin soaked in
10% acetic acid. The iodothyronines are adsorbed on the resin,
and the inorganic iodide passes through and can be quantified
directly. Others have used precipitation of horse serum (car-
rier) with TCA to separate the free iodide (143). Either method
requires determination that equimolar amounts of
125
I and
[
125
I]T
3
(or [
125
I]T
2
if [
125
I]rT
3
is used as a substrate) are pro-
duced, particularly if the level of enzyme activity is low or if
non-deiodinative pathways are present. This is achieved by
separating the [
125
I]-iodothyronines and
125
I by paper des-
cending chromatography, HPLC, or UPLC. For calculating
the 5¢D activity, the percentage of product generated is mul-
tiplied by 2 because the specific activities of the labeled
products are only half that of the substrate. D1 activity is
generally expressed as picomoles or femtomoles of product
per unit of time per milligram of protein (144–147).
104 BIANCO ET AL.
Assay for D2 activity. This assay is carried out as described
for the D1 assay with the following modifications: the pre-
ferred substrate is T
4
(0.5–2 nM) and if tissues that also contain
D1 activity are being used, 1 mM PTU is included in the in-
cubation medium. If D3 is present, it should be saturated by
adding an appropriate excess of T
3
(146–149).
Assay for D3 activity. The preferred substrate for D3 is T
3
(0.5–2 nM), and since T
3
is a very poor substrate for D1 and
D2, it can be successfully used to assay D3 activity in tissues
that also contain D1 and/or D2 activity. Addition of EDTA to
the assay buffer is not necessary. In this assay, since the ra-
dioactive substrates are labeled in the outer-ring only, the
iodide generated cannot be detected, and thus quantitation of
the reaction necessitates the separation of the T
2
product from
the substrate T
3
by HPLC or descending paper chromatog-
raphy. Since the T
2
generated is also a good substrate for D3, it
is important to dilute the enzyme preparation such that the
percent deiodination is relatively low (<20%) and T
2
is the
only detectable compound produced (147,150–152).
&
RECOMMENDATION 13b
Studies of deiodinase activity in cell or tissue preparations
containing more than one deiodinase are feasible but spe-
cial provisions must be considered so that each deiodinase
activity is measured independently.
Commentary. Often times not only the substrate of a
given deiodinase but also the deiodination product can serve
as a substrate for another deiodinase co-expressed in the same
cell or tissue. Thus, to observe the activity of a specific deio-
dinase requires the use of enzyme-specific inhibitors and
different substrate types and/or concentrations (Fig. 9) (153–
156). For example, measuring outer-ring D1 activity in the
presence of D2 is possible if T
4
or rT
3
are used at their D1 Km
(i.e., approximately 1 lM and 0.5 lM, respectively), which
saturates D2. Measuring inner-ring D1 activity in the presence
of D3 is possible if T
3
or T
3
sulfate (T
3
S) is used at its D1 Km
(i.e., approximately 5 lM), which saturates D3. D1 is effi-
ciently inhibited in the presence of 1 mM PTU, which has only
minimal effect on D2 activity. This eliminates D1 activity from
any D2 or D3 activities, which are not affected by PTU.
No specific inhibitors are known for D2 or D3. However,
both D2 and D3 activities can be eliminated by adding a rel-
ative excess of their preferred substrate (i.e., T
4
or T
3
, respec-
tively), saturating the enzyme and outcompeting the
radioactive tracer. For example, measuring D2 activity in the
presence of D3 is possible by using rT
3
or T
4
as substrate and
adding 100–1000 nM T
3
in the reaction mixture. Adding the T
3
excess saturates D3 binding sites but does not interfere with
the D2 activity. However, a high purity reagent is preferred
because, at high concentrations, even small amounts of a
contaminant iodothyronine will interfere with the assay.
&
RECOMMENDATION 14a
The Vmax of any of the deiodinases measured under op-
timum conditions of substrate and cofactor availability can
be used as a surrogate for the amount of functional enzyme
present in a cell or tissue at any given moment.
Commentary. While D1 and D3 are expressed at levels
that can be measured by Western blotting or immunocyto/
histochemistry using commercially available antisera
(65,157,158), the combined low expression of D2 with the un-
availability of sufficiently high affinity D2 antisera has im-
paired quantification of D2 protein (159). In fact, no consensus
exists as to the validity of anti-D2 antiserum/antibodies for
measurement of endogenous protein; results need to be
evaluated on a case-by-case basis. However, an estimation of
D2 protein levels can be obtained reliably by measuring D2
Vmax, which reflects the maximal amount of active enzyme in
a cell or tissue. D1 and D3 can be assessed either via Vmax or
Western blot. The protein levels of all three deiodinases can be
quantified after labeling of deiodinase-expressing cells with
75
Se, immunoprecipitation, and resolution by sodium dodecyl
FIG. 9. Strategies to measure individualized deiodinase activity in the presence of other deiodinases. (A) Two strategies to
assess type II deiodinase (D2) activity in the presence of type I deiodinase (D1; e.g., in the human thyroid): (i) use 1 mM 6-n-
propyl-2-thiouracil (PTU) to inhibit D1 or (ii) use 1–2 nM [
125
I]T
4
in the presence or absence of 100 nM cold thyroxine (T
4
). (B)
To measure D2 activity in D2/D3 co-expressing tissues (e.g., the brain), use 1000 nM cold T
3
to saturate D3. (C) To measure
D2 activity in D1/D2/D3 co-expressing tissues (e.g., rodent gonads, placenta, cerebrum, and skin), inhibit D1 with 1 mM
PTU and saturate D3 with 1000 nM T
3
.
INVESTIGATING THYROID HORMONE ECONOMY AND ACTION 105
sulfate polyacrylamide gel electrophoresis, but this takes
several weeks per assay for D2 (93). Note that biological al-
teration of Km (e.g., amino acid mutations) could alter the
relationship between Vmax and protein levels.
&
RECOMMENDATION 14b
The expression of the deiodinase genes can be determined
in cells or tissues by measuring the respective mRNA
levels.
Commentary. Measuring D1 mRNA levels is straight-
forward, and given its exquisite responsiveness to T
3
, it con-
stitutes a sensitive marker of peripheral thyroid status in
rodents (72). D2 is a cAMP-dependent gene and its mRNA
levels can be up-regulated several-fold during sympathetic
stimulation of BAT (161). At the same time, D2 exhibits strong
posttranslational regulation via ubiquitination and proteaso-
mal degradation by different components of the endoplasmic
reticulum (162–165). Thus, mRNA levels do not necessarily
reflect protein expression or enzyme activity (162,166,167).
The Dio3 gene is intronless, and thus appropriate controls are
required given that even minimal DNA contamination could
affect the RT-qPCR results. D3 has a relatively long half-life,
and thus a high level of D3 activity can persist even after an
elevation in D3 mRNA has dissipated (168,169). In situ hy-
bridization is particularly useful in the brain given the com-
plexity in the expression patterns of D2 and D3 (170–172) (see
Sections I.2 and J.1 for technical considerations).
[D.2] Deiodination in intact cells
Background. Deiodination can be studied in intact live
cells. Essentially, established cell lines or primary cultures
expressing 5¢D and/or 5D activity are exposed to the appro-
priate [
125
I]-substrate (as previously discussed) and the radi-
olabeled products determined in the media and/or in cell
sonicates as already described.
&
RECOMMENDATION 15
T
4
and T
3
metabolism via deiodination can be measured in
live cells with the advantage that studies are performed
with physiological cofactor.
Commentary. Cells are incubated with [
125
I]T
4
or
[
125
I]T
3
in the presence of media containing 0.1%–1.0% BSA
with added T
4
and/or T
3
to yield a physiological concen-
tration of free hormone in the low picomolar range (Table 2).
Metabolites in the media or in the cells are separated by
liquid or paper chromatography (as already described). The
reaction time is typically 24 hours or less. Color-free culture
media must b e use d due to interfe ren ce wit h th e ass ay. Th e
desired concentration of the substrate is achieved by incu-
bating cells in serum-free media containing 0.1% BSA and
including the appropriate concentration of nonradioactive
T
4
and/or T
3
. Time courses can be established by sampling
the media followed by quantification o f the pro ducts as
described above. Res ults are express ed as picomoles o r
femtomoles of product per unit of time; correction for the
number of cells can be achieved by cell counting or by de-
termining the DNA or protein content. Studies in cells
containing more than one deiodinase are feasible and
should include appropriate controls in which one of the
deiodinative pathways is saturated with an excess of non-
labeled substrate (136,137,152).
Coculture systems with more than one cell type have been
developed in which thyroid hormone transport, metabolism,
and action can be studied simultaneously (i.e., D2-expressing
H4 human glioma cells and D3-expressing SK-N-AS human
neuroblastoma cells; Fig. 10) (173). Such a system has been
used to demonstrate that paracrine signaling by glial cell-
derived T
3
activates neuronal gene expression.
[D.3] Deiodination in perfused organs
Background. Outer-ring and inner-ring deiodination can
be studied in perfused organs such as kidney and liver. Es-
sentially, freshly harvested organs expressing 5¢D and/or 5D
activities are perfused with buffered solutions containing BSA
and radiolabeled or cold iodothyronine substrates. The per-
fusate is collected over time and analyzed for deiodination
products using immunoassays or HPLC. This approach is
advantageous because it allows for the study of tissue-specific
deiodinative pathways under physiological or defined patho-
physiological conditions.
&
RECOMMENDATION 16a
T
4
to T
3
conversion and the urinary iodothyronine excre-
tion can be studied in preparations of perfused rat kidney
or liver.
Commentary. In preparations of kidney perfused with T
4
in BSA-containing buffer there is a liner increase in T
3
pro-
duction with the increase in the perfusate FT
4
concentration
that determines tissue T
4
uptake. FT
4
levels can be adjusted by
increasing or decreasing the concentration of total T
4
or the
BSA concentration. Addition of PTU decreases renal T
3
pro-
duction by about 60% without affecting tissue T
4
uptake, il-
lustrating the presence of D1. In this setting there is no net
renal rT
3
production from T
4
, and degradation and urinary
excretion of T
3
are negligible (174–176).
Similar to the kidney, the perfused rat liver readily extracts
T
4
from perfusion medium and converts it to T
3
. Production of
T
3
is a function of the size of the liver, the uptake of T
4
, and
Table 2. Free Fractions of T
4
and T
3
( ·100
and Expressed as Percentage) for the Common
Media Types Containing Different Percentages
of Bovine Serum Albumin/Serum
Media type T
3
(%) T
4
(%) Reference
BSA
a
4% 0.56 0.09 (648)
1% 2.18–2.3 0.29–0.32 (745,746)
0.5% 3.45–3.78 0.41 (747–749)
0.1% 3.39–3.6 (747,748)
Bovine serum
10%
b
4 0.45 (316)
10% (stripped)
c
1.5–4 0.8 (750,751)
10% (Tx) 0.4–2.0 (281)
a
Direct measurements by equilibrium dialysis.
b
Direct measurement by ultrafiltration.
c
Calculated based on total and free hormone concentrations
(equilibrium dialysis).
T
4
, thyroxine; BSA, bovine serum albumin; Tx, thyroidectomized.
106 BIANCO ET AL.
level of D1 expression. Production of T
3
, but not T
4
uptake, is
decreased by PTU (177,178).
&
RECOMMENDATION 16b
Placental inner-ring deiodination can be studied in situ
using a guinea pig perfusion system.
Commentary. Whereas placental D2 and D3 activities can
be measured in cell dispersions or tissue sonicates (179–182),
in situ preparations can be used to study the placental D3
pathway (183,184). In an anesthetized pregnant guinea pig, the
placenta is surgically exposed and a single umbilical artery and
the umbilical vein cannulated, while the fetus is removed. The
fetal side of the placenta is perfused through the umbilical ar-
tery with buffered solution containing BSA and [
125
I]T
3
.Pla-
centa effluent fractions are collected at timed intervals (up to 2
hours) from the umbilical vein cannula. The contents of the
perfusion buffer and the various effluent fractions are then
separated and analyzed by HPLC for their iodothyronine
content (183). In this setting outer-ring deiodination of T
4
or rT
3
is minimal (184) possibly because placental outer-ring deiodi-
nation is greatest in the zone immediately adjacent to the
uterine wall (185), distant from the fetal side of the organ.
[D.4] Deiodination in whole animals
Background. The study of the deiodination pathways in
the whole animal is challenging since one is dealing with three
deiodinases, each of which can deiodinate not only T
4
and/or
T
3
, but also the products of these reactions. In addition there are
no known agents that will selectively and completely inhibit
individual deiodinases. Thus, until mice deficient in one or
more of the deiodinases became available, it was very difficult
to investigate the role of the individual deiodinases in vivo.
&
RECOMMENDATION 17
Total body deiodination can be studied in live animals
following administration of radiolabeled iodothyronines
(e.g., [
125
I]T
4
or [
125
I]T
3
). Iodothyronines can be injected
acutely or long-term via a mini-pump. This approach
FIG. 10. In vitro modeling of thyroid hormone deiodination and transport in the brain. (A) Schematic representation of the
Transwell System in which an insert is placed on a six-well plate and cells (D2-expressing H4 glial cells) are seeded inside the
insert; D3-expressing neuroblastoma cells (SK-N-AS) are seeded at the bottom of the six-well plate. After cells are seeded, both cell
types are kept separated overnight and then placed together in the same multiwell plate as indicated. (B) At the end of the
incubation medium samples are collected, extracted and processed through a UPLC or HPLC connected to the flow gamma
counter to separate and quantify the activity of each iodothyronine. The red arrow indicates the pathway completed by the
column eluate through the gamma counter; courtesy Dr. Antonio Bianco. (C) Chromatograms of H4 cell medium at the indicated
times after addition of
125
I-T
4
. Typical peaks of
125
-T
3
and
125
I are shown after 24 hours. (D) Same as in (C),exceptthat
125
I-T
3
was
addedtoculturesofSK-N-AScellsand
125
I-T
2
and
125
I-T
1
peaks are visualized. (E) Same as in (C),exceptthat
125
I-T
4
was added to
H4 and SK-N-AS cocultures and the indicated peaks are visualized. UPLC, ultrahigh performance liquid chromatography; HPLC,
high performance liquid chrom atograph y; T
1
,5¢-monoiodothyronine. Reproduced with permission from Freitas et al. (173).
INVESTIGATING THYROID HORMONE ECONOMY AND ACTION 107
allows for studies to be conducted under controlled phys-
iologic conditions. Additional information may be ob-
tained by administering more than one iodothyronine
labeled with different isotopes of iodine.
Commentary. Total body phenolic-ring deiodination can
be readily assessed in rodents following the administration of
a single dose of [
125
I]T
4
or [
125
I]T
3
. The rodents are then placed
in individual cages that permit the separate collection of urine
and feces. After an appropriate period of time (42–72 hours)
the fraction of the injected radioactivity excreted in the urine
can be determined, and the vast majority of this will be in-
organic iodide (186). This protocol can be used to assess the
effects of different conditions such as fasting, cold exposure,
hyper- or hypothyroidism, and the absence of one or more of
the deiodinases on total deiodination. This type of study can
be carried out using a relatively low level of radioactivity.
However, obtaining any reliable information regarding the
extent of T
4
conversion to T
3
or rT
3
by this method is com-
plicated by the fact that both compounds are cleared from
serum at a faster rate than T
4
itself.
Some information concerning the labeled products
formed from [
125
I]T
4
or [
125
I]T
3
can be obtained using ani-
mals implanted with osmotic mi ni-pumps containing the
labeled hormones. Once the daily excretion of radioactivity
in the urine and fec es becomes constant, indicating that the
animals have reached isotopic equil ibrium, the animals can
be sacrificed. The serum, tissues, urine, and f eces can be
obtained an d the identi ty of their labeled compounds de-
termined, following their extraction and separation by
HPLC or chromatography. Parallel studies using mice defi-
cient in one or more of the deiodinases will shed light on the
relative importance of the different metabolic pathways.
Several extraction procedures have been published and care
must be taken to determine and correct for extraction effi-
ciency (83,187).
Results of these studies must be interpreted carefully and
alternative possibilities considered. For example, it is well-
established that both T
4
and T
3
can be conjugated with sulfate
in vivo, and these conjugates are excellent substrates for the
inner-ring deiodinating activity of the D1. The thyroid hor-
mones are also conjugated with glucuronic acid and most of
the T
4
in the kidney is present as T
4
glucuronide, which is not
detected in an RIA (137). In addition, a significant fraction of
the metabolites generated from T
4
and T
3
, including the glu-
curonide conjugates, are to be found in the bile, the intestinal
contents, and feces.
[D.5] Non-deiodination pathways
of thyroid hormone metabolism
Background. Nonnutrient substances that reach the gas-
trointestinal system are also known as xenobiotic compounds
and are identified in the liver through specific receptors; for
example, pregnane X receptor (PXR) and constitutive an-
drostane receptor (CAR). Binding to these xenobiotic sensing
receptors, PXR and/or CAR, induces the expression of met-
abolic enzymes including Phase I (the cytochrome P450
family of enzymes) and Phase II (e.g., sulfotransferases, glu-
curonosyltransferases, and glutathione-S transferases). These
pathways modify xenobiotic molecules by introducing reac-
tive or polar groups into their molecules to increase their
water solubility, thereby facilitating their elimination. The
Phase II pathways include glucuronidation, methylation,
sulfation, acetylation, glutathione conjugation, and amino
acid conjugation.
Perhaps because of the high iodine content, the iodothyr-
onines are also metabolized by some of the Phase II pathways;
that is, conjugation of the phenolic hydroxyl group with sul-
fate or glucuronic acid (188). In fact, a relatively small portion
of the daily thyroid hormone production is processed through
these pathways, but it can be much more if these metabolic
pathways are induced by xenobiotic compounds such as
central nervous system (CNS)-acting drugs (e.g., phenobar-
bital, benzodiazepines), calcium channel blockers (e.g.,
nicardipine, bepridil), steroids (e.g., spironolactone), reti-
noids, chlorinated hydrocarbons (e.g., chlordane, dicholoro-
diphenyltrichloro-ethane, tetrachlorodibenzo-p-dioxin, and
polyhalogenated biphenyls such as polychlorinated biphenyl
and polybrominated biphenyl, among others) (189).
Sulfated iodothyronines do not bind to the TRs, and sulfa-
tion mediates the rapid and irreversible deiodination of io-
dothyronines by D1. Therefore, the concentrations of sulfated
iodothyronines in serum are normally low. Inner-ring deiodi-
nation (inactivation) of T
4
and T
3
by D1 is markedly facili-
tated after sulfation, whereas outer ring deiodination of T
4
is
blocked after sulfation. As expected, the D1 KO mouse exhibits
marked increase in fecal excretion of [
125
I]-iodothyronines
during the 48 h after injection of [
125
I]T
4
or [
125
I]T
3
,whereas
urinary excretion of [
125
I]iodide was markedly diminished
(186). Notably, D2 and D3 are not capable of deiodinating
sulfated iodothyronines. Plasma levels and biliary excretion of
iodothyronine sulfates are increased in fetal and cord blood,
nonthyroidal illness (NTI), fasting, and by inhibition of D1
activity with PTU or iopanoic acid (190). Under these condi-
tions, T
3
S may function as a reservoir of inactive hormone from
which active T
3
may be recovered by action of tissue sulfatases
and bacterial sulfatases in the intestine (188).
Iodothyronine glucuronides are rapidly excreted in the bile.
However, this is not an irreversible pathway of hormone
disposal. After hydrolysis of the glucuronides by bacterial
b-glucuronidases in the intestine, part of the liberated io-
dothyronines are reabsorbed, resulting in an enterohepatic
cycle of iodothyronines (188). Nevertheless, about 20% of
daily T
4
production appears in the feces, probably through
biliary excretion of glucuronide conjugates. Glucuronidation
is catalyzed by UDP-glucuronyltransferases (UGTs) that uti-
lize UDP-glucuronic acid (UDPGA) as the cofactor. UGTs are
localized in the endoplasmic reticulum predominantly of
liver, kidney, and intestine. Most UGTs are members of the
UGT1A and UGT2B families (191).
In general, the relation between tissue enzyme activities
for the different iodothyronines and the expression of indi-
vidual isoenzymes is hardly known, especially for the sulfo-
transfereases (SULTs). Many SULTs exhibit overlapping
substrate specificities. In addition, multiple SULTs in the same
tissue can be involved in the sulfoconjugation of the same
iodothyronines, resulting in clear redundancy [see Wu et al.
(192) for an overview]. As a consequence, the biochemical
properties of tissue SULT activity reflect the composite effect
of different isoenzymes. Expression or protein levels of the
different isoenzymes can be studied using RT-qPCR and/or
Western blotting, but do not necessarily reflect the overall
tissue sulfoconjugation activities (192,193).
108 BIANCO ET AL.
&
RECOMMENDATION 18a
Sulfotransferase activities and their intrinsic properties can
be determined in cell or tissue preparations using 3¢-
phosphoadenosine-5¢-phosphosulfate (PAPS) as the sulfate
donor.
Commentary. Iodothyronine sulfation is catalyzed by
SULTs using PAPS as the sulfate donor. In humans, SULTs
show overlapping substrate specificity. In humans, they can
be subdivided into different families, SULT1, SULT2, SULT4,
and SULT6. All members of the human SULT1 family (i.e.,
hSULT1A1, -1A2, -1A3, -1B1, -1C2, -1C4, and -1E1) catalyze
the sulfation of iodothyronines (194). hSULT1A1-3, -1B1, and -
1C4 and also native enzymes in liver have a substrate pref-
erence for 3,3¢-diiodothyronine (T
2
), which is catalyzed much
faster than T
3
and rT
3
, whereas T
4
sulfation is negligible.
hSULT1E1 equally prefers T
2
and rT
3
over T
3
and T
4
but is the
only known SULT so far having significant T
4
-sulfotransfer-
ase activity (194). As a consequence, T
2
is generally used to
study sulfotransferase activities in tissues homogenates under
different conditions. Different human SULTs have also been
shown to catalyze the sulfation of iodothyronamines (195).
Sulfotransferase activity assay. There are multiple accept-
able protocols for sulfotransferase assays. It is best to use
radiolabeled iodothyronines with the highest available spe-
cific activity and assays should contain background controls
and samples running in parallel in tubes without PAPS.
Due to relatively high Km values, it may be difficult to out-
compete the radiolabeled iodothyronine by a large excess of
substrate. SULT activity can be analyzed by incubation of
0.1 lM[
125
I]-iodothyronine of interest for 30 minutes with
tissue homogenate, cytosol, or recombinant SULT in the
presence or absence (blank) of 50 lM PAPS in PBS-EDTA
buffer (194). The reaction is stopped by the addition of 0.1 M
HCl and the products separated by filtration in Sephadex LH-
20 minicolumns into iodide, sulfated iodothyronines, and
nonsulfated iodothyronines (196).
&
RECOMMENDATION 18b
Iodothyronine glucuronidation activities are determined in
microsomal cell or tissue preparations using UDPGA as a
cofactor.
Commentary. Glucuronidation of T
4
and T
3
is catalyzed
by different members of the UGT1A family (197). Usually, this
involves the glucuronidation of the hydroxyl group, but
human UGT1A3 also catalyzes the glucuronidation of the
side-chain carboxyl group, with formation of so-called acyl
glucuronides. Interestingly, Tetrac and TRIAC are much more
rapidly glucuronidated in human liver than T
4
and T
3
, and
this occurs predominantly by acyl glucuronidation (198). Acyl
glucuronides are reactive compounds that may form covalent
complexes with proteins. It is unknown if this is a significant
route for the formation of covalent iodothyronine–protein
complexes.
Glucuronidation activity assay. There are multiple accept-
able protocols for glucuronyl-transferase assays. Radio-
labeled iodothyronines with the highest available specific
activity should be used and assays should contain back-
ground controls. Iodothyronine glucuronidation activity can
be analyzed by incubation of 1 lM of the [
125
I]-iodothyronine
of interest for 60 minutes with microsomes in magnesium
chloride–containing Tris-HCl (pH 7.8) buffer, in the presence
or absence (blank) of 5 mM UDPGA (197,199). When tissue
microsomes are analyzed, 1 mM PTU may be added to the
reaction mixture to prevent iodothyronine deiodination
without affecting their glucuronidation. Reaction is stopped
by the addition of ice-cold ethanol and glucuronide formation
is analyzed in supernatant by chromatography on Sephadex
LH-20 minicolumns as already described.
[E] Inducing Hypothyroidism
and Thyroid Hormone Replacement
Overview. Hypothyroidism is a pathological state in
which thyroid hormone signaling is decreased systemically or
locally in one or more tissues. As a result of the depletion
of nuclear T
3
, there is modification in the expression of T
3
-
responsive genes, decreasing the biological effects of thyroid
hormone. Induction of hypothyroidism has been used tradi-
tionally to define and characterize T
3
-responsive processes, an
approach that can be used in animals or in cultured cells. In
rodents this is accomplished by decreasing serum levels of T
3
and in cultured cells by reducing the free T
3
concentration in
the medium, below physiological levels. Alternatively, TR
antagonists have been developed and used in cells.
[E.1] Hypothyroidism in animals
Background. Serum T
3
concentrations in rodents can be
reduced surgically by total thyroidectomy, or medically
by treatment with antithyroid drugs or
131
I. In addition, there
are a number of rodent strains in which key genes in the
hypothalamic-pituitary-thyroid (HPT) axis exhibit spontane-
ous mutations or have been genetically modified, ultimately
disrupting thyroid hormone synthesis and/or secretion. The
experimental approach for achieving hypothyroidism should
take into consideration the age of the animals (i.e., prenatal,
early postnatal, after weaning). Lastly, disruption and/
induction of deiodinases or thyroid hormone transporters
may result in tissue-specific hypothyroidism.
&
RECOMMENDATION 19
Body weight gain should be monitored during induction of
hypothyroidism. For cross-reference between experiments,
an observed plateau in body weight gain should be taken to
define a state of systemic hypothyroidism. T
3
-responsive
gene expression and enzyme activities, particularly liver
D1 expression or activity, can be used as additional mea-
sures of thyroid status.
Commentary. A drop in serum T
4
and an elevation in
serum TSH are the first indications of a disruption in the
function of the HPT axis. More time is usually needed to re-
duce serum T
3
. Given that T
3
is the biologically active thyroid
hormone, in theory, a state of systemic hypothyroidism could
be considered to exist after serum concentrations of T
3
have
dropped below the normal range. However, given that the
drop in serum T
3
might not reflect tissue T
3
concentration at
early time points, most studies define systemic hypothyroid-
ism to exist when a major thyroid hormone–dependent bio-
logical effect is observed; that is, body weight gain plateaus,
or the expression level of a T
3
-responsive gene or the activity
of a T
3
-responsive enzyme is reduced. While in human
studies, TSH would be an ideal parameter to define thyroid
INVESTIGATING THYROID HORMONE ECONOMY AND ACTION 109
status, in rodents normative data are lacking to allow for
cross-study comparison using TSH as a primary marker of
thyroid status.
Because rodents continue to grow throughout life and be-
cause growth hormone secretion is exquisitely sensitive to T
3
in rats, growth, as assessed by body weight gain, is a very
sensitive marker of thyroid hormone action in rats (200,201).
An approximately 100 g rat stops putting on weight about 3
weeks after total thyroidectomy; at this time point, thyroid
hormone levels have dropped by 75% and pituitary growth
hormone content is almost undetectable (201). An absence of
growth for 2 weeks has been proposed as the gold-standard
for defining a state of severe systemic hypothyroidism (202);
shorter periods of time (e.g., 5 days of documented growth
plateau) should be acceptable as well. Notably, arrest of linear
growth (as determined by tail length), which is very sensitive
to thyroid hormone, would be an excellent and more gener-
ally applicable indicator of hypothyroidism, although pub-
lished data specifically documenting this are lacking. Less
data are available for mice, in which growth hormone is less
T
3
-sensitive, so a plateau in body weight gain is not a reliable
indicator of hypothyroidism.
A molecular approach to define intermediate states of
systemic hypothyroidism would involve assessment of the
expression of T
3
-responsive genes or the activity of T
3
-
responsive enzymes such as cardiac mRNA levels of myosin
heavy chain (MHC) isoforms or hepatic a-glycerophosphate
dehydrogenase (a-GPD) activity or D1 activity. While any T
3
-
responsive tissue could be examined, traditionally the liver
has been the organ of choice to assess systemic hypothy-
roidism given its high number of TRs and well-defined T
3
-
responsive pathways. Liver D1 activity is considered the most
sensitive genetic index of systemic thyroid status and thus can
assist in the characterization of very subtle states of disruption
in thyroid hormone signaling (72) (see Sections D.1 and I.2).
&
RECOMMENDATION 20a
Systemic hypothyroidism can be induced by surgical total
thyroidectomy in adult rats or mice. Systemic hypothy-
roidism is achieved usually between 5 and 8 weeks after
surgery. Hemi-thyroidectomized rats develop a mild form
of systemic hypothyroidism.
Commentary. Surgical total- or hemi-thyroidectomy in
rodents is a widely used procedure given the ready access to
the thyroid gland (Fig. 11) (203). In general, when performing
total thyroidectomy on a 100 g rat or a 20 g mouse, the para-
thyroid glands are preserved. Because surgical skills and
parathyroid anatomy may vary, some investigators assume
that the procedure will result in postsurgical hypoparathy-
roidism, and thus provide animals with a solution of 2%–4%
calcium lactate in 5% dextrose ad libitum as the only fluid
source for at least 10 days postsurgery.
Thyroparathyroidectomized rats and mice can also be
purchased from commercial vendors.
&
RECOMMENDATION 20b
Systemic hypothyroidism can be induced by chemical
thyroidectomy caused by the administration of antithyroid
drugs. The time frame of hypothyroidism onset is variable
and depends on the type of drug and regimen used.
Commentary. Chemical inhibition of the thyroid gland
can be induced via administration of MMI, PTU, KClO
4
,or
NaClO
4
. These drugs can be given through daily intraperi-
toneal injections (e.g., PTU 1–2 mg/100 g BW, MMI 1–5 mg/
100 g BW (204)), added to the chow (e.g., 0.02%–0.15% PTU or
0.01% MMI or KClO
4
1.25% (205)) or the drinking water
(0.01%–0.1% MMI, 0.01%–0.1% PTU, or 0.1%–1% KClO
4
or
NaClO
4
). A major pitfall of this strategy is that all of these
antithyroid drugs have a bitter taste and, when added to the
FIG. 11. Dissection of the rodent thyroid gland. (A) Surgical incision, (B) isolate the salivary glands, (C) dissociate the
muscles, (D) free the trachea, (E) section the lateral muscles, (F) slide a needle underneath the trachea, revealing the thyroid
gland. Modified with permission from a web posting by Prof. Emeritus Jean-Pierre Herveg and Christian Regaert.
110 BIANCO ET AL.
drinking water, consumption should be monitored. In fact,
PTU is a commonly used bitter stimulus in studies of taste
physiology (206). Empirically, it has been found that KClO
4
is
less palatable than NaClO
4
. Because taste sensitivity varies
among inbred mice, water consumption will be variably af-
fected, and in some strains significant dehydration may occur.
At the same time, drug consumption may also vary because
rats and mice may stop drinking bitter fluids and can with-
stand total water deprivation for several days before hyper-
osmolality develops (207). Sucrose can be added to the water
to prevent this, but responses are variable (208). Thus, if
drinking water is the delivery method chosen, body weight
and drinking consumption should be monitored carefully and
drug concentrations adjusted appropriately. A minor con-
sideration in rodents is that treatment with PTU or MMI ele-
vates serum calcitonin concentration and, with prolonged use,
C-cells will exhibit physiologic or reactive hyperplasia. This is
not related to the state of systemic hypothyroidism; the cause
is not known.
Serum T
3
levels will fall below normal within 10–15 days
depending on the drug concentration and route used. As an
example, there was a 20-fold increase in serum TSH in rats
treated for up to 3 weeks with 0.05% MMI in drinking water
(209). Treatment with 0.1% PTU in drinking water, which
also inhibits peripheral conversion of T
4
to T
3
, resulted in a
faster reduction in serum T
3
and increase in serum TSH than
in rats treated with 0.1% MMI. However, these differences
were no longer observed after 3 weeks of treatment (209).
A state of hypothyroidism can be achieved faster if PTU or
MMI are combined with 1% KClO
4
in the drinking water, a
potent NIS inhibitor (210,211). However, their combination
will also magnify the effect that these drugs have on water
consumption.
A reasonable alternative is to use 0.15% PTU-containing
low-iodine food pellets, which is commercially available and
has been used successfully (212). It has been reported that a
systemic state of hypothyroidism can be achieved after 4
weeks on this diet; a faster approach includes combining
water 0.01% MMI, which is a concentration that does not seem
to interfere with water consumption. In long-term experi-
ments (months), compensatory TSH-driven thyroid growth
can raise serum T
4
back to nearly normal levels, even while
animals are on the PTU–low-iodine diet. Thus, it is important
to document a state of systemic hypothyroidism throughout
the experiment.
&
RECOMMENDATION 20c
Systemic hypothyroidism can be induced by radioactive
iodide ablation of the gland. In adult animals thyroid
hormone release is markedly reduced after 1 week of
131
I
administration. Growth in newborn animals plateaus after
3–4 weeks of
131
I administration.
Commentary. Intraperitoneal administration of
131
Iin
solution results in transient or permanent hypothyroidism in
adult rats (250 g BW) depending on the dose used (213). A
progressive thyroidal degenerative process is observed after
administration of 300 or 525 lCi
131
I but regeneration occurs
after several months; regeneration is not seen when 875 lCi is
used (213). Other tissues may be damaged by the radiation:
fibrosis of the parathyroid glands has been observed when
525 or 875 lCi of
131
I is used. There can be tracheal damage
and transient renal effects (213). No damage to the liver has
been observed (213). The earliest metabolic indicator of de-
creased glandular function is a drop in the protein bound
131
I,
which is observed as early as 24 hours after the injection of
875 lCi of
131
I (214). This approach is also effective in newborn
rats after intraperitoneal administration of 80 lCi of
131
I (215).
In a 20–25 g mouse, direct ablation can be achieved with
200 lCi of
131
I (216).
&
RECOMMENDATION 20d
A state of systemic hypothyroidism can be achieved
quickly and effectively by the use of combined approaches.
Commentary. Any combination of low-iodine diet, anti-
thyroid drugs, surgery, and radioactive iodide could be used
quickly and effectively to achieve a state of severe systemic
hypothyroidism. Typically,
131
I is preceded by a low-iodine
diet. Alternatively, surgical thyroidectomy can also be fol-
lowed by a low-iodine diet for 2 weeks, followed then by
remnant ablation with
131
I; the postsurgical ablation dose for a
200–250 g rat can be *80 lCi (217); for mice, 25–50 lCi has
been used successfully.
&
RECOMMENDATION 21
Secondary hypothyroidism can be induced in rats by sur-
gical hypophysectomy.
Commentary. Surgical hypophysectomy in rodents re-
quires skill and experience. The parapharyngeal approach to
remove the pituitary gland in rodents has been used tradi-
tionally (218,219). This approach has been largely supplanted
by the transauricular approach (220,221). This procedure
leaves the animal deaf in one or both ears, which would affect
their performance in certain behavioral tests. Typically, an
anesthetized 50–300 g rat is immobilized in a Hoffman–Reiter
hypophysectomy instrument, which allows for the stereo-
tactic placement of the tip of a suction needle proximate to the
pituitary gland, which is then aspirated. The animal is then
removed from the instrument and provided postoperative
care. The completeness of hypophysectomy is confirmed by
the absence of weight gain over a 4 week period after surgery.
Perioperative care of these animals is critical: hypophysecto-
mized animals have difficulty regulating body temperature
and thus require constant environmental conditions; a war-
mer room (temperature between 72F and 76F) is ideal. This
procedure increases diuresis and thus cages containing con-
tact bedding must be changed frequently. Given their meta-
bolic fragility, experiments are typical conducted after the first
6 to 10 postoperative days. All hypophysectomized animals
are maintained on drinking water containing appropriate
electrolites and 5% glucose, ad libitum. Water supplements
commence on the day of surgery and should continue for 14
days. To produce animals that exhibit selective secondary
hypothyroidism, hypophysectomized animals are treated
with daily replacement doses of growth hormone, luteinizing
hormone, follicle stimulating hormone, prolactin, testosterone
propionate, 17b-estradiol benzoate, cortisone, or vehicle by
subcutaneous injections (222).
&
RECOMMENDATION 22a
The Nkx2 homeobox 1 (Nkx2.1; T/ebp, TTF-1) KO mouse is
a model of thyroid dysgenesis.
INVESTIGATING THYROID HORMONE ECONOMY AND ACTION 111
Commentary. The Nkx2.1 KO mouse lends itself for
studies of fetal thyroid gland development as well as for
studies in which fetal thyroidal secretion does not exist.
Thyroid-specific enhancer-binding protein (T/ebp)/Nkx2.1-
null mouse thyroids degenerate by embryonic day (E) 12–13
through apoptosis, whereas T/ebp/Nkx2.1-heterogyzgous
mice exhibit hypothyroidism with elevated TSH levels. The
Nkx2.1 (T/ebp, TTF-1) transcription factor regulates thyroid-
specific gene transcription but is also important for develop-
ment of pituitary, lung, and ventral brain regions (223). The
homozygous Nkx2.1 KO mouse is not suitable for postnatal
studies since these animals die at birth due to profoundly
hypoplastic lungs, a severely defective hypothalamus, and
absence of thyroid and pituitary glands. On the other hand,
Nkx2.1-heterozygous mice were shown to exhibit hypo-
thyroidism with elevated TSH levels and have a neurological
defect, although they appear to be healthy and are fertile. This
thyroid phenotype is caused by reduced expression of the
TSH receptor due to T/ebp haplo-insufficiency.
&
RECOMMENDATION 22b
The Pax8 KO mouse, the TSH receptor KO mouse, and the
mouse homozygous for the autosomal recessive gene hy-
pothyroid (hyt) exhibit severe congenital hypothyroidism
of fetal onset.
Commentary. Given that the Pax8 gene governs thyroid-
specific transcription, its inactivation results in a small thyroid
gland that lacks follicle formation (224). Pax8 KO mice have
undetectable serum T
4
and T
3
levels, increased postnatal
mortality, and growth retardation. In contrast to the Nkx2.1
KO, the Pax8 KO mouse survives birth but dies after weaning.
Postnatal treatment with T
4
or T
3
, will rescue these animals.
Withdrawal of treatment during adult life produces a severely
hypothyroid animal that can survive for 6 months or longer.
The molecular defect in the hyt mouse is an inactivating
point mutation in the gene encoding the TSH receptor
(TSHR). Neonatal hyt/hyt mice have reduced serum T
4
ranging from 1/5 to 1/6 of normal as well as significantly
delayed somatic and behavioral development. The hyt/hyt
mouse provides an ideal model for exploring the effect of
severe primary inherited hypothyroidism in utero and in the
early neonatal period (225,226). The TSHR KO mouse is
profoundly hypothyroid with no detectable thyroid hormone
in serum and elevated serum TSH, exhibiting developmental
and growth delay as well as infertility if not supplemented
with T
4
(227).
&
RECOMMENDATION 22c
The mouse with a missense mutation in the dual oxidase 2
(DUOX2) gene and the Cog mouse (congenital goiter) are
less severe models of congenital hypothyroidism.
Commentary. Mice with a missense mutation in the
DUOX2 gene exhibit a milder form of congenital hypothy-
roidism. The DUOX2 gene is involved in the generation of
H
2
O
2
for thyroid peroxidase, the enzyme that catalyzes iodine
organification into thyroglobulin for thyroid hormone syn-
thesis. A valine to glycine replacement (V674G) in DUOX2
explains the thyroid dyshormonogenesis (228). The homozy-
gous DUOX2 mutant develops to adulthood but exhibits
dwarfism and suffers from hearing impairment. Serum T
4
levels are 10% of wild-type levels and are associated with an
approximately 100-fold elevation in serum TSH and a dys-
plastic anterior pituitary. This mouse model of congenital
hypothyroidism can be made more severe by simultaneous
targeting of DUOXA1 and DUOXA2, which are specific
maturation factors required for targeting of functional DUOX
enzymes to the cell surface. Homozygous males and fe-
males exhibit goiter and congenital hypothyroidism with
undetectable serum T
4
and 500- to 2500-fold elevated TSH
levels, respectively (45).
The cog mouse exhibits an autosomal recessive mutation
that has been mapped to the thyroglobulin gene (230,231).
Young adult cog mice exhibit reduced rate of growth, mild
anemia, lower serum T
4
and T
3
, and elevated serum TSH.
Thyroids from mutant mice are hypertrophied, deficient in
colloid, accumulate less iodine (partially susceptible to per-
chlorate discharge), and have a marked deficiency in thyro-
globulin content (230).
&
RECOMMENDATION 22d
The TRH KO mouse is a model of tertiary hypothyroidism,
in which the disruption in thyroid function is only mild.
Commentary. The TRH KO mice are viable and fertile
and exhibit normal development ( 80). However, they have
a marked decrease in serum T
4
(about 60% of control val-
ues), elevation of serum TSH level, and diminished TSH
biological activity. The an terior pituitary has decreased
TSH immune-positive cells, which is corrected by TRH but
not thyroid hormone replacement. The TRH KO mouse
exhibits a slight growth delay by 4 weeks of age that is
normalized by 8 weeks of age or by treatment with T
4
(80).
Because of the role played by TRH in the pancreatic islets,
the TRH KO mice also exhibit hyperglycemia, which is
accompanied by impaired insulin secretion in response to
glucose.
&
RECOMMENDATION 23a
Cell- and tissue-specific forms of hypothyroidism related to
deiodinase activities can be studied in vivo via disruption of
the D2 pathway or induction of the D3 pathway.
Commentary. The activity of the deiodinases can modify
T
3
levels in a cell-specific fashion without affecting circulating
thyroid hormone levels (84). A disruption in the D2 pathway
has been shown to decrease T
3
production locally and disrupt
thyroid hormone signaling in D2-expressing cells (92,173,232–
234). This is illustrated by the approximately 50% reduction in
T
3
content in the D2 KO brain (187). D2 inhibitors can be used
but with caution due to off-target effects. For example, inhi-
bition of D2 has been identified as the underlying cause in the
elevated serum TSH associated with amiodarone treatment
(235), but amiodarone also inhibits D1 and can have its own
TR effects as well. Similar concerns exist with iopanoic acid,
used to inhibit D2 and disrupt thyroid hormone signaling in
BAT (236,237). The use of the D2 KO animals constitutes the
best model to study the localized hypothyroidism due to
disruption of the D2 pathway.
Tissues expressing D3 have decreased thyroid hor-
mone signaling as a result of rapid T
3
inactivation to T
2
and thus constitute a model of localized hypothyroidism
(62,63,238,239). This can be exaggerated during D3
112 BIANCO ET AL.
reactivation during illness, ischemia, or hypoxia that leads to
localized disruption in thyroid hormone signaling in different
tissues such as myocardium and brain (62–65,238,241).
&
RECOMMENDATION 23b
Cell- and tissue-specific hypothyroidism related to iodo-
thyronine transport can be studied in vivo via disruption of
various transporter systems.
Commentary. Inactivating mutations in the gene encod-
ing MCT8 disrupt T
3
transport across plasma membrane and
decrease thyroid hormone signaling (107,108). Given that D2
is expressed in glial cells and TR mostly in neurons, the MCT8
inactivating mutation is thought to promote neuronal hypo-
thyroidism by disrupting the paracrine effects of glia-made T
3
at a critical time for CNS development. In fact, the existence of
this paracrine mechanism was verified in cocultures of glia
and neuronal cell lines (173). Clinical evidence supporting this
mechanism was first obtained in patients with severe neuro-
logical phenotype exhibiting the Allan-Herndon-Dudley
syndrome (107,108).
[E.2] Thyroid hormone replacement in animals
Background. The molar ratio of T
4
to T
3
in rat thyroids is
8:1 (242) and the estimated T
4
/T
3
ratio in the thyroidal se-
cretion is 5:1 (1000 pmol/d of T
4
to 190 pmol/d of T
3
), indi-
cating a small contribution of thyroidal T
4
-to-T
3
conversion to
the daily T
3
production in the rat (132). Thus, the prohormone
T
4
is the major secreted iodothyronine in iodine-sufficient rats.
While all T
4
is secreted by the thyroid, about 60% of daily T
3
production is via peripheral deiodination of T
4
, and the re-
maining 40% secreted directly from the thyroid gland. No-
tably, the human thyroid gland contributes with only about
20% of the daily T
3
production (132).
In theory, thyroid hormone replacement in rodents could
be modeled on the protocols applied in humans; that is,
starting thyroid hormone replacement based on weight and
monitoring serum TSH and T
4
. However, in rats L-T
4
mono-
therapy cannot simultaneously normalize serum T
3
,T
4
, and
TSH, presumably because of the higher thyroidal T
3
produc-
tion in rodents (243,244). Thus to normalize T
3
and TSH, the
serum T
4
must be higher than normal in rats, and the same
applies to tissue T
3
and T
4
content. Historically, investigators
have used the daily T
4
production rate, about 10 ng/g BW, as
a replacement dose in L-T
4
monotherapy in rats because this
dose is sufficient to normalize the growth rate (200). However,
if normalization of serum TSH is desired, a higher dose (ap-
proximately 20 ng/g BW) is required. While a similar ap-
proach for L-T
4
treatment has been used in hypothyroid mice
(i.e., treatment dose being equivalent to the daily production
rate) normative data are lacking to determine whether serum
thyroid function tests and tissue T
3
and T
4
levels respond
similarly as in rats.
&
RECOMMENDATION 24a
The daily T
4
replacement dose in adult rats previously
rendered hypothyroid should approximate the daily pro-
duction rate if growth rate is held as the primary endpoint
defining the euthyroid state. Higher doses of L-T
4
must be
used if serum TSH is to be normalized. If other thyroid
hormone–sensitive endpoints are to be considered, the
appropriate T
4
dosing must be determined empirically on a
case-by-case basis.
Commentary. In general, normalization of serum T
3
or
biological parameters such as growth, the expression of a T
3
-
responsive gene, or the activity of a T
3
-responsive enzyme in
the liver are acceptable parameters when considering thyroid
hormone replacement in rats previously rendered hypothy-
roid. Daily delivery of *8 ng/g BW to adult rats previously
rendered hypothyroid has been shown to normalize serum T
3
levels, growth rate, and the activity of liver a-GPD, a very
sensitive index of systemic thyroid status in rats (202).
For practical reasons, administration of T
4
is the method
used most frequently to restore euthyroidism (200,202,245,
246). However, because of the substan tial direct thyroidal
secretion of T
3
in ro dents, the ideal thyroid hormone re-
placement regimen would feature a m ixture of T
4
and T
3
.
Some studies have established the optimal dosing for com-
bination therapy in rodents (244). When T
4
is used alone for
replacement, tissue T
3
content can vary across a large
number of tissues because of the role played by the deiodi-
nases as a local source of tissue T
3
(via conversion from T
4
)
(243,244).
The calculated daily T
3
production is approximately
2.7 ng/g BW (132). Notably, a slightly lower dose of T
3
,
2.0 ng/g BW, has been shown to restore growth and the ac-
tivity of liver a-GPD in previously rendered hypothyroid rats
(202). However, administration of T
3
alone is not sufficient to
restore euthyroidism in all tissues, particularly in those with a
significant contribution of the D2 pathway (243,244).
&
RECOMMENDATION 24b
The parenteral route is preferred to deliver thyroid hor-
mone to rodents; for example, intraperitoneal, subcutane-
ous, or osmotic pumps or subcutaneously implanted
pellets.
Commentary. T
4
and T
3
solutions are prepared in 40 mM
sodium hydroxide and diluted in saline for injections. When
protected from light, stock solutions can be stored at - 20Cfor
a few weeks. Daily injections bypass concerns about variable
intake if the hormone is administered in food or water (see
following text). However, it should be borne in mind that in-
jections produce a rapid, supraphysiological peak in systemic
T
4
or T
3
levels. Alternative delivery methods for T
4
or T
3
ad-
ministration such as subcutaneous pumps or pellets may be
preferred if relatively constant rates of delivery are desired. T
4
and T
3
can be mixed with water or food, but these methods
carry the intrinsic variability of food or water consumption that
might preclude their use in some experiments. Treatment du-
ration can span days, weeks, or months. Preferably, doses
should be divided daily (every 12 hours) given the relatively
short half-lives of T
3
and T
4
in rodents (2 and 8 hours, respec-
tively). Blood sampling for measurement of thyroid hormone
levels must take into account the timing of the last injection.
&
RECOMMENDATION 24c
Age-appropriate regimens for T
4
replacement should be
used for hypothyroid neonatal mice.
Commentary. In models of congenital or neonatal hypo-
thyroidism, T
4
replacement should be started at birth (day 0)
INVESTIGATING THYROID HORMONE ECONOMY AND ACTION 113
and continued on a daily basis through at least postnatal day
10, a critical thyroid hormone–dependent developmental
period in mice (247). Serum T
4
levels in severely hypothyroid
mice can be restored to age-matched control levels via sub-
cutaneous injections of T
4
on days 0–5 (4 ng/g), 6–8 (5.8 ng/g),
and 9–10 (9.1 ng/g) for the mouse (248–251). Injections can be
performed using a 0.5 mL syringe with a 30-gauge saline-
treated needle.
&
RECOMMENDATION 24d
Mouse embryonic stem cells can be driven to differentiate
into functional thyroid follicular cells in vitro, restoring
systemic euthyroidism when transplanted into hypo-
thyroid mice.
Commentary. Transient overexpression of the transcrip-
tion factors NKx2-1 (formerly called TTF-1) and Pax8 directs
mouse embryonic stem cells to differentiate into thyroid fol-
licular cells. When treated with TSH these cells organize into
3D follicular structures and activate thyroid functional genes
including NIS, TSH, thyroglobulin, and thyroperoxidase
genes (252,253). These in vitro derived follicles show iodide
organification activity and when grafted in vivo into athyroid
mice are able to restore serum thyroid hormone levels and
promoted subsequent symptomatic recovery (252).
[E.3] Hypothyroidism in cultured cells
Background. Culture medium generally contains animal
serum, and so cultured cells are exposed to iodothyronine
concentrations that depend on the source of the serum. For
example, if fetal bovine serum (FBS) is used to support cell
growth, the ambient T
3
and T
4
concentrations in the medium
will reflect the thyroid hormone levels of the calves used to
produce the serum. Historically, serum taken from thyroid-
ectomized animals or animals treated with antithyroid drugs
have been used to generate hypothyroid media (254). Because
of a lack of commercial availability of such serum, chemical
stripping has become more common, either with charcoal,
anion exchange AG1-X8 resins, or both (255,256). Depending
on the cell type being studied, using serum-free media may
also be an option, though most cells propagate better in media
containing serum.
&
RECOMMENDATION 25a
Cells in culture can be made hypothyroid using media
supplemented with either (i) charcoal-stripped serum,
(ii) resin-stripped serum, (iii) serum obtained from hypo-
thyroid animals, or (iv) medium not supplemented with
serum.
Commentary. Using ‘defined’ media to induce hypo-
thyroidism has a number of caveats. While many reports in-
dicate > 99%’ of T
3
and T
4
are removed via standard
stripping protocols, some variability exists in the degree of T
3
and T
4
removal achieved by the various procedures. This
variability may arise both from factors related to the type of
serum and from the method of stripping (257,258). Further-
more, current chemical stripping methods do not allow for the
depletion of thyroid hormone without also depleting other
circulating factors. Many small molecules such as growth
factors or hormones (including sex steroids, adrenal steroids,
and vitamin D) may also be removed during stripping, and
thus the biologic changes seen over time cannot be described
as representing isolated hypothyroidism (257). Experimental
design should compare groups with thyroid hormone re-
placed (versus vehicle replaced) to enable specific attribution
of the results to thyroid hormone deficiency. If a more isolated
depletion of T
3
and T
4
is necessary, preparation of hypothy-
roid animal serum has been shown to be potentially cost-
effective, despite an initial appearance of impracticality (254).
Using depleted serum has been shown to induce cellular
hypothyroidism, as assessed by the expression level of T
3
-
responsive genes, in as little as 24 hours (259). In theory, the
time needed to achieve hypothyroidism could be shortened
by frequently changing the media (e.g., every few hours in-
stead of daily); this would accelerate the depletion of intra-
cellular T
3
stores. Similarly, prolonged culture time in T
3
- and
T
4
-depleted media would be expected to induce a more pro-
found state of cellular hypothyroidism. Regardless of the
approach, the extent of hypothyroidism for each experimental
condition should be determined via measurement of T
3
-
responsive endpoints (e.g., expression of T
3
-responsive genes
or T
3
-responsive enzymes) to facilitate interassay comparison.
&
RECOMMENDATION 25b
Cells in culture can be made hypothyroid using a TR
antagonist.
Commentary. NH-3 is a TR antagonist that strongly in-
hibits transcriptional activation by T
3
(260–265). Treatment
with this compound has been shown to block thyroid hor-
mone dependent processes such as spontaneous Xenopus
laevis tadpole metamorphosis. In theory, this drug could be
used to induce hypothyroidism in other vertebrates, but this
remains to be established experimentally.
[F] Increasing Thyroid Hormone Signaling
Overview. Classically, thyrotoxicosis is thought of as a
pathophysiological state in which thyroid hormone signaling
is increased systemically (i.e., throughout all tissues of the
organism) as a result of increased T
3
binding with its nuclear
receptors. Experimentally, systemic thyrotoxicosis can be
modeled in vivo via treatment of animals with thyroid hor-
mone. In vitro, cells can be treated with media supplemented
with thyroid hormone above the concentrations seen in blood.
However, in the latter case, the effects of cross-talk between
thyrotoxic tissues (i.e., the indirect, interactive effects medi-
ated via second messengers, are absent); the state of ‘direct
thyrotoxicosis’ created is thus distinct from systemic thyro-
toxicosis. ‘Tissue-specific increase in thyroid hormone sig-
naling’ is a more recent concept, arising as the result of local
deiodinase activity that increase nuclear T
3
concentration of
certain tissues without necessarily altering plasma thyroid
hormone concentration (66).
[F.1] Thyrotoxicosis in animals
Background. Administration of thyroid hormone to an
otherwise euthyroid rodent leads to systemic thyrotoxicosis,
the intensity of which is directly related to the magnitude of
the elevation in serum T
3
concentration. A number of ap-
proaches have been used to achieve systemic thyrotoxicosis,
including intraperitoneal injection of thyroid hormone,
114 BIANCO ET AL.
addition to chow or drinking water, and subcutaneous pellet
or mini-pump insertion.
&
RECOMMENDATION 26a
Acute systemic thyrotoxicosis can be induced by parenteral
administration (intraperitoneal or intravenous) of 1.0 lg
T
3
/g BW, a TR-saturating dose of T
3
. Genomic effects can
be seen as early as 60 minutes and physiologic effects
starting at about 6 hours.
&
RECOMMENDATION 26b
Long-term systemic thyrotoxicosis can be induced by
chronic treatment with T
3
or T
4
over days, weeks, or
months. Routes of administration include parenteral or
supplementation of food or water. For chronic studies, TR-
saturating doses should be avoided due to cachexia and
death. To allow comparisons between studies, T
3
doses
should be given as multiples of the daily T
3
or T
4
replace-
ment dose.
Commentary. The half-life of T
3
in rodents following in-
jection has been measured to be approximately 4 hours, and
for T
4
about 11 hours (132). Given these short half-lives, in
either acute or chronic experiments, sampling for measure-
ment of T
3
or T
4
serum levels must take into account the
timing of the last injection, in particular for T
3
. Furthermore
multiple injections should be considered if biological effects
are to be measured after several half-lives have passed. After
multiple injections with classical doses of T
3
(100 lg/100 g
BW), the intense state of thyrotoxicosis triggers loss of body
weight and increases mortality, such that experimental
treatment duration should not extend beyond 4–5 days. Ty-
pical doses of T
4
or T
3
well tolerated in long-term experiments
(2–3 months) are 10- to 25-fold the daily production rate.
While food- and water-based drug delivery methods are
subject to feeding variability, this may not be critical for
studies of chronic thyrotoxicosis; for example, powdered ro-
dent diet containing 3 mg of T
4
and 1 mg of T
3
per kilogram
has been used successfully to induce chronic thyrotoxicosis
(205). Consideration should be given to more consistent
methods such as subcutaneous pumps or pellets (243,244).
Depending on the endpoint to be studied, the choice of
iodothyronine is important, since T
4
is converted to T
3
via
D1 and D2. Graves’ patients can sometimes exhibit a T
3
-
predominant form of thyrotoxicosis, whereas thyroid cancer
patients are generally treated with pharmacologic doses of T
4
to achieve TSH suppression. That being said, in studies not
focusing on the roles of the deiodinases, T
3
is preferred be-
cause it does not require further activation and thus elimi-
nates the activating deiodinases as a variable.
Classically, large doses of T
3
have been used in rodent ex-
periments, designed to saturate the nuclear T
3
receptors rap-
idly and thus maximize the phenotypic events studied as
endpoints, while minimizing experimental time and thus cost.
For example, a typical dose for rats or mice would be 100 lg/
100 g BW, producing measurable genomic and physiologic
changes within a few hours (266). It should be noted that the
doses of T
3
classically used for rodent experiments are clearly
in the pharmacologic range and are much higher than the
pathophysiologic range seen in nature. Some authors ques-
tion whether such high doses may produce off-target artifacts
or overly enhance nongenomic effects of T
3
.
&
RECOMMENDATION 27
In vivo activation of the D2 pathway or inactivation of the
D3 pathway can be used to study cell- and tissue-specific
increase in thyroid hormone signaling.
Commentary. cAMP-induced D2 activation in BAT leads
to rapid saturation of TR with locally generated T
3
and in-
duction of T
3
-responsive genes and the activity of T
3
-
responsive enzymes, without affecting circulating thyroid
hormone levels (66,86,233,237,267–269,271). This is also ob-
served in animal models of transgenic Dio2 expression in the
heart (273,274). In contrast, D3 inactivation results in localized
increase in thyroid hormone signaling as evidenced in the D3
KO mouse (64,239,275–279).
[F.2] Thyrotoxicosis in cultured cells
Background. For cell culture–based experiments, tissue-
specific or more properly cell type–specific thyrotoxicosis can
be modeled via addition of T
3
to the medium. Medium thy-
roid hormone concentrations can be determined directly, but
in most cases they are estimated based on published values
for the free fractions in a given serum.
&
RECOMMENDATION 28a
Acute thyrotoxicosis can be achieved in cell culture via
addition of T
3
in the media. The free fraction of T
3
multi-
plied by the total concentration of T
3
gives an estimate of
the free hormone concentration. If free fractions are not
determined directly, estimates of free hormone concentra-
tion can be made based on published values of the free
fractions.
Commentary. If exact thyroid hormone concentrations
within cell culture media must be kno wn, then direct mea-
surements would be required. In most cases, however,
rough estimates are thought to suffice (Table 2). Notably,
thyroid hormones adhere to plasticware, so it is important
to add a suitable carrier protein in order to model the bound
and free thyroid hormone fractions i n cell culture. Stripped
serum or fatty acid–free BSA are typically used, with T
3
(and/or T
4
)replaced;thesamecaveatsaboutlossofother
hormones still apply when this experimental approach is
used.
Few studies have reported equilibrium dialysis-based
measurement of free fractions in stripped serum, but one
study found that for FBS it is similar to the estimated range of
unstripped FBS (0.4%–4% for T
3
) (281). BSA with fatty acids
may have T
3
, thus fatty acid–free BSA should be used (282). It
should also be noted that most commercially available forms
of T
4
have at least trace T
3
contamination that must be con-
sidered. Importantly, recall that the free fraction of T
3
(or T
4
)
increases as the percentage of serum in the media decreases.
Thus, the free hormone concentration stays relatively constant
as the percentage of serum decreases.
As in the case with rodents, pharmacologic dosing of T
3
has
been used historically, with high concentrations designed to
achieve saturation of nuclear receptors rapidly. For example,
100 nM total T
3
in 0.5% BSA would give an estimated free T
3
concentration of around 3500 pM; for comparison, the free T
3
concentration of euthyroid human serum would be closer to
3–8 pM (about 1000 times lower). Even severely thyrotoxic
INVESTIGATING THYROID HORMONE ECONOMY AND ACTION 115
patients would not be expected to achieve such high free T
3
concentrations, only reaching values of about 20 pM (283), or
as high as 35 pM in the case of an accidental ingestion of a
massive dose of L-T
4
(284). Some investigators have sug-
gested that pharmacologic dosing can artificially enhance
nongenomic and hypothetical off-target effects of T
3
; if these
are not a concern, the dose can be maximized for effect. Ide-
ally, T
3
doses that are within or near the physiological range
should be used.
&
RECOMMENDATION 28b
While inducing thyrotoxicosis in cell cultures, the possible
presence of deiodinases should be considered for each cell
type being studied.
Commentary. Knowing the dei odina se ac tivitie s of a
particular cell line may be important experimentally. Unless
the effects of deiodination are of particular interest, thyro-
toxicosis should be induced via addition of T
3
only, not T
4
.
This is because D1 activates T
4
(converting it to T
3
), such
that ce lls treated with T
4
will have variable media T
3
con-
centrations depending on the level of D1 activit y and the
volume of media and frequency of media changes. This ef-
fect is expected to be more intense in D2-expressing cells,
given its much higher c atalytic efficiency. In contrast, D3
activity would lower the media concentration o f T
3
(con-
verting it to T
2
), increasing the dosing requirement to sustain
thyrotoxicosis.
[F.3] Use of thyroid hormone analogues
Background. The existence of two TR isoforms (i.e., TRa
and TRb) indicates that different signaling pathways (and
perhaps sets of biological effects) are downstream of each one
of these molecules. This is further strengthened by the ob-
servation that the distribution of TR isoforms is heteroge-
neous among different tissues/cells (285,286). For example,
bone is a predominantly TRa tissue while liver is a predomi-
nantly TRb tissue; thus the rationale for developing molecules
that exhibit TR-isoform specificity.
Two caveats should be considered while selecting the
concentration and dose of these molecules:
(i) Pharmacokinetic data: TR selectivity depends on ligand
concentration and selective tissue uptake. For example,
eprotirome appears to have only modest selectivity for
TRb in vitro, yet its profound effects to lower lipids and
cholesterol result from selective availability to the liver,
possibly due to selective uptake. There is no information
on transporter selectivity for different ligands or whether
these ligands use the same or different transporters
compared to T
3
and T
4
. The use of relatively high doses or
high media concentrations minimizes or eliminates TR
selectivity. There is currently no or only limited com-
prehensive pharmacokinetic studies for most analogues,
making it difficult to define concentration and doses in
comparative experiments with T
3
or other analogues.
(ii) Ligand transport into the cell and cell nucleus: the
mechanisms leading to cellular uptake and nuclear
concentration of the analogues have not been defined.
Thus, even when using equimolar media concentra-
tions of two such molecules their concentration in the
nucleus, around the TR, could be different. This could
enhance or eliminate any biological advantage of TR
selectivity exhibited in vitro.
&
RECOMMENDATION 29
Thyroid hormone signaling can be triggered in vivo and
in vitro by thyroid hormone analogues, some of which have
tissue selectivity or selectivity for TR isoforms (i.e., TRa1,
TRb1, TRb2). A number of highly selective TRb analogues
target tissues exhibiting predominance of TRb (e.g., liver
and pituitary gland). The utilization of equimolar doses
and concentrations of different analogues and T
3
is re-
commended as a starting point in comparative studies.
Commentary. 3,5-Diiodothyropropionic acid (DITPA) is
a carboxylic acid thyroid hormone analogue that binds TRa
and TRb with nearly identical affinities (287). Studies in rats
and rabbits indicate that DITPA has positive inotropic effects
but minimal chronotropic and metabolic effects outside the
cardiovascular system (287,288). Its combined use with cap-
topril improved ventricular performance and reduced end-
diastolic pressure in the rat postinfarction model of heart
failure (289). Additional effects included attenuation of the
acute inflammatory response and reduction of myocardial
infarct size (290), and improvement of maximal perfusion
potential of the hypertrophied myocardium surviving a
myocardial infarction (291). However, other studies reported
side effects (292,293) or failed to obtain positive results with
DITPA in similar settings (294,295).
Tiratricol is an acetic acid thyroid hormone analogue that
exhibits about 3.5-fold greater in vitro affinity for TRb and 1.5-
fold greater affinity for TRa compared to T
3
, with an ap-
proximately threefold selectivity for TRb (296,297). However,
there is only moderate TRb selectivity in cell culture studies
(298). In rats, tiratricol has been shown to have antidepressant
effects (299), thermogenic effects in the BAT (300,301), and
some degree of organ specificity due to its enhanced liver and
skeleton effects and reduced cardiac effects (302). The use of
tiratricol in a number of settings, including patients with TSH
hypersecretion and athyrotic patients, reduced serum TSH
(303–305) and lowered serum cholesterol levels without af-
fecting heart rate but elevated biochemical markers of bone
turnover (306,307).
A new generation of highly selective TRb agonists has
much less affinity for TRa while preserving affinity for TRb
(308). The molecules in the GC family (i.e., GC-1 and GC-24)
are noniodinated compounds that were designed based on
experimental data obtained for the TR structure (309). The use
of these molecules has suggested the possible existence of a
‘therapeutic window’ through which predominantly TRb-
mediated biological effects (e.g., lowering serum cholesterol
and acceleration of energy expenditure) can be triggered with
relatively little activation TRa-dependent pathways (310–
316). However, it is uncertain whether tissue specificity in vivo
is due to selectivity for TR binding or tissue concentration of
these molecules or both. For example, the TRb selective ago-
nist GC-1 concentrates preferentially in the liver as opposed to
heart, skeletal muscle, or brain (310). The KB family of mole-
cules (i.e., KB-141 and KB-2115) has been used successfully in
animals (317–320) and in hypercholesterolemic patients kept
on statins to lower serum cholesterol even further while
sparing the heart and bone (321,322). In these patients there
were only minimal alterations in thyroid function tests.
116 BIANCO ET AL.
However, clinical studies with these molecules were sus-
pended due to undesirable cartilage side effects observed in
dogs (323).
The KB general structure was subsequently used as a
scaffold to design a family of indane (hydrocarbon com-
pounds) derivatives that exhibit potent and selective thyro-
mimetic activity (324). KTA-439, a representative indane
derivative, displays the same high human TRb selectivity in a
binding assay as KB2115 and higher liver selectivity in a
cholesterol-fed rat model (324).
An alternative strategy to obtain tissue selectivity is to use
molecules that concentrate in specific tissues by virtue of
undergoing local activation such as the cytochrome P450 ac-
tivation of a prodrug that is a phosphonate-containing TR
agonist. This molecule exhibits increased TR activation in the
liver relative to extrahepatic tissues and an improved thera-
peutic index (319). MB07811 undergoes first-pass hepatic ex-
traction and cleavage, generating the TR agonist MB07344
that distributes poorly into most tissues and is rapidly elimi-
nated in the bile (319). MB07811 lowers serum cholesterol in
hypercholesterolemic rats, rabbits, monkeys, and humans
beyond what was achieved with statins alone (325), is supe-
rior to a TRb-selective agonist in the diet-induced obese
mouse model (319), and reduces hepatic steatosis in rats and
mice models (326).
Selective analogues for TRa have also been developed (327)
and shown to be effective in promoting TRa-dependent neu-
rogenesis in X. laevis (328). However, TRa selectivity was lost
when the same compound was tested in rats, with CO23 ac-
tivating thyroid hormone–responsive genes in liver and heart
(329).
[G] Iodine Deficiency and Maternal–Fetal Transfer
of Thyroid Hormone
Overview. Iodine is the major constituent of thyroid
hormone. Simply pu t, the most activ e form of thyroid h or-
mone, T
3
, can be viewed as three atoms of iodine attached to
a phenoxyphenyl scaffold. The unique spatial positioning of
the iodine atom s confers high affinity for the TR, a ligand-
dependent transcriptional regulator. It is remarkable that
vertebrates evolved to have a scarce environmental element,
iodine, play such an important role in embryogenesis,
growth, metabolism, cognition, and adaptation to disease
states. The sea is the main source of io dine and perhaps
the major biological role played by iodine reflects the idea
that life began to exist and evolved in the ocean. Iodine-
containing clouds are formed over the oceans and blown
inland, with rain depositing iodine over the land. Depending
on the type of soil, iodine is retained for some time or quickly
washed to the rivers and back to the oceans. Consequently,
theiodinecontentofplants,crops,andanimalsinanyspe-
cific geographical region depends on the iodine content of
the s oil. About 1.5 billion people live in geographic areas
of iodin e ins ufficiency, requiring s ome form of iodine sup-
plementation to prevent the consequences of perinatal
hypothyroidism.
Iodide is absorbed in the small intestine and avidly taken
up by the thyroid via NIS located in the basal-lateral mem-
brane of all thyrocytes. Once inside the cells, iodide diffuses
towards the relatively positively charged lumen of the thyroid
follicle, exiting the thyrocyte via pendrin channels located in
the apical membrane. Adjacent to microvilli of the apical
membrane, iodide is oxidized and conjugated to specific ty-
rosine residues in the thyroglobulin molecule, a process cat-
alyzed by thyroid peroxidase. Subsequently, iodinated
thyroglobulin molecules are reabsorbed via pinocytosis and
digested in lysosomes, releasing T
4
,T
3
, and small amounts of
thyroglobulin as well as other iodinated molecules into the
circulation.
Iodine deficiency can pose a serious threat to the thyroidal
capacity to synthesize thyroid hormones. It is generally ac-
cepted that limited iodine availability acts as an evolutionary
pressure that favors the development of compensatory
mechanisms, which minimize the impact of iodine deficiency
on thyroid economy. Deciphering these mechanisms is critical
not only for our understanding of thyroid gland function and
thyroid hormone economy but also to formulate strategies
that can be used to treat and prevent the irreversible conse-
quences of fetal and neonatal hypothyroidism resulting from
iodine deficiency. The most severe condition is neurological
cretinism due to severe iodine deficiency and hypothyroxi-
nemia during the first half of pregnancy resulting in irre-
versible brain damage (330). Thus, experimental models of
iodine deficiency have been widely used to analyze the
adaptive mechanisms developed in animal models and the
impact of different degrees of iodine deficiency on thyroid
economy.
[G.1] Iodine deficiency in rodents
Background. The main consequences of iodine deficiency
are goiter, hypothyroxinemia (242), increased NIS expression
(331), increased monoiodotyrosine (MIT)/diiodotyrosine (DIT)
and T
3
/T
4
ratios within the thyroglobulin (332,333), prefer-
ential T
3
synthesis and secretion by the thyroid (242,332), in-
creased D1 and D2 activities in the thyroid and D2 activity in
BAT (334,335), low T
4
in plasma and tissues (71,336,337), and
a TSH concentration that is slightly elevated, together with a
normal or slightly decreased T
3
level in serum and several
tissues (71). Tissue uptake of T
4
and T
3
increases. Only some of
the endpoints of thyroid hormone action are affected but to a
lesser degree than in overt hypothyroidism (71,338). The
chow diet fed to rodents in accredited animal facilities con-
tains enough iodine (0.4–1 lg/g) to allow for a normal daily
iodine intake (*5–10 lg/d) and thyroid function. Two strat-
egies can be used, independently or in combination, to pro-
mote iodine deficiency: (i) feeding with a low iodine diet
(<0.02 lg/g) that reduces iodine intake (to 0.2–0.4 lg/d) or (ii)
treatment with drugs that inhibit NIS and thus thyroidal io-
dine uptake.
&
RECOMMENDATION 30a
A state of iodine deficiency can be achieved by feeding
rodents with an LID (containing <0.02 lg iodine/g). Milder
degrees of iodine deficiency can be achieved with less
stringent diets. Effects on thyroid economy can be seen as
early as 10 days, but for most parameters clear effects require
at least 1 month of LID.
&
RECOMMENDATION 30b
Iodine deficiency in animals can be documented by moni-
toring urinary iodine excretion, which should be about 5- to
10-fold lower in the LID animals.
INVESTIGATING THYROID HORMONE ECONOMY AND ACTION 117
&
RECOMMENDATION 30c
Intrathyroidal iodine deficiency can be achieved via inhi-
bition of NIS following administration of 1% KClO
4
in the
drinking water. Thyroidal iodine stores can be depleted in 1
week using this approach. Longer treatments will lead to
systemic hypothyroidism, which must be taken into account.
Because KClO
4
is bitter, caveats apply for this strategy as
discussed previously (see Recommendation 19b).
&
RECOMMENDATION 31a
The impact of iodine-deficiency on the thyroid gland itself
can be assessed by measuring gland size (at least a twofold
increase should be seen) and histologically, with LID
leading to varying degrees of hyperplasia and hypertro-
phy. The amount of colloid in the follicles is reduced, and
the epithelial cells lining the narrowed follicular spaces
consist of columnar cells instead of normal cuboidal cells.
&
RECOMMENDATION 31b
The impact of iodine deficiency on thyroid economy can be
monitored by measuring serum T
4
, and the T
3
/T
4
ratio in
thyroglobulin. TSH elevation is not an early finding. Si-
milarly, serum T
3
is preserved until iodine deficiency is
severe.
Commentary. LID is commercially available from differ-
ent vendors, but checking the iodine content of the diet is
strongly recommended given that it is not unusual to find
higher than reported iodine content. Remington-type diet has
been used successfully; synthetic diets containing casein or
other proteins that are a source of iodine should be avoided. In
accredited animal facilities, rodents drink deionized distilled
water, which does not containing significant amounts of
iodine.
In all settings, control animals should also be fed LID but
drink water containing KI ad libitum to provide about 5–10 lg
iodine/day. The combination of LID and KClO
4
results in
severe iodine deficiency, a reduction in serum T
3
, and overt
hypothyroidism. When KClO
4
is employed it should be used
in low percentages ( £ 0.005%) to avoid undesirable environ-
mental pollution (71,339).
&
RECOMMENDATION 32
Gestational or neonatal iodine deficiency can be achieved
by feeding dams with the LID. Given the short duration of
gestation in rodents, dams should be pretreated with LID
so that hypothyroxinemia is achieved before the onset of
pregnancy.
Commentary. This strategy leads to profound fetal hypo-
thyroidism because maternal transfer of T
4
is markedly di-
minished, and T
4
is the only source of T
3
for the fetal brain
(340,341). After birth, maternal T
4
is concentrated in the milk
and given to the pups prior to weaning, thus resulting in
amelioration of their hypothyroidism (341).
Severe iodine deficiency before and throughout gestation
results in a rat model of neurological cretinism (330) with brain
alterations similar to those described in human affected pop-
ulations (342–344). Given that feeding on LID can potentially
decrease fertility, particularly if the state of iodine deficiency is
severe enough to reduce serum T
3
, LID alone might not act
quickly enough to create a state of iodine deficiency in the fetus.
In this case, KClO
4
or its combination with LID can be used
successfully to promote fetal iodine deficiency.
[G.2] Placental transfer of thyroid hormone
Background. The placenta functions as an interface be-
tween the maternal and fetal circulations, allowing for the
exchange of nutrients, gases, and many other types of mole-
cules. Thyroid hormones are not only metabolized by pla-
cental cells via D2 and D3, but there is also a net flux of thyroid
hormones transport to the fetus, which ensures the presence
of T
4
and T
3
in the fetal circulation before the fetal thyroid is
fully developed. Thyroid hormones transport can be studied
in the rat by inhibiting the fetal thyroid with MMI given to the
mother. In this setting, different levels of both T
4
and T
3
are
found in fetal tissues.
&
RECOMMENDATION 33
Placental transfer of thyroid hormones from mother to
fetus can be studied in MMI-treated and thyroid hormone–
replaced pregnant dams by measuring T
4
and T
3
content in
fetal tissues.
Commentary. Pregnant dams are given 0.02% MMI in the
drinking water, starting on the 14th day of gestation. MMI
crosses the placenta and has been shown to inhibit fetal thy-
roidal function and promote severe fetal hypothyroidism
with elevated serum TSH and low thyroid hormone levels in
blood and tissues. Mothers are treated with thyroid hormone
(T
4
,T
3
, or both) and, at any time during pregnancy or im-
mediately after delivery, fetuses are dissected and blood/
tissues obtained for extraction and determination of T
4
and T
3
contents (345,346). Administration of tracer [
125
I]T
4
to preg-
nant dams and its detection in fetal tissues has also been used
to study placental transfer of thyroid hormone (347).
[H] Models of Nonthyroidal Illness
Overview. Thyroid economy is markedly affected by ill-
ness, fasting, or other major life-threatening conditions. This is
known as NTI syndrome, euthyroid sick syndrome, or low T
3
syndrome (348). NTI may be viewed as part of the acute phase
response to illness or injury, a defense mechanism predomi-
nantly mediated by cytokines (349). A number of animal and
cell models have been developed to study the pathogenesis
and pathophysiologic consequences of NTI, including fasting,
injury or illness, and lipopolysaccharide (LPS) administration.
In general, during NTI there is a multilevel suppression of
the HPT axis and decreased thyroidal secretion (350). A drop
in serum leptin levels plays a central role in the fasting-
induced changes in thyroid economy via neuropeptide Y-
mediated TRH suppression (351), with leptin administration
restoring the fasting induced state of central hypothyroidism
(352). There are also major modifications in the metabolic
pathways of thyroid hormone, such as accelerated extra-
thyroidal inactivation of thyroid hormone via deiodination
(353), glucuronidation, and sulfation (351); it is not yet clear
whether extrathyroidal T
3
production is decreased (354).
However, it is well accepted that central hypothyroidism is
the main driving force behind the changes in circulating
thyroid hormone levels and thyroid economy associated with
fasting or starvation. Also important is understanding that
118 BIANCO ET AL.
changes in thyroid economy during NTI may be very different
in the acute versus a more chronic phase (355). Defining these
mechanisms and the pathophysiological implications of these
changes is central to our understanding of NTI, which affects
the majority of patients admitted to any general hospital.
&
RECOMMENDATION 34
Fasting is widely used as a model of NTI to study its impact
on thyroid hormone production, metabolism, and action.
Commentary. Fasting promotes central hypothyroidism,
reducing serum levels of thyroid hormone without a corre-
sponding elevation in serum TSH. Mice and rats can be fasted
for periods of hours or days depending on standards and
protocols set by the local institutional animal committee.
Usually adult rats are fasted no longer than 2–3 days and
adult mice 1 day. In rats, fasting is associated with an ap-
proximately 50%–75% drop in serum T
4
and T
3
that takes
place in the first 48 hours (356). In mice, the drop in serum T
3
is
about 50% and takes place by 36 hours of fasting (357). Fasting
also leads to a reduction in liver D1 and increase in D3 ac-
tivities (358,359). However, the reduction in serum T
3
ob-
served in the fasted rat is mostly secondary to a reduction in
thyroidal secretion (360).
&
RECOMMENDATION 35a
Major bodily insults or severe illness in rodents such as
extensive surgery, burns, inflammatory pain, bacterial in-
fection, or prolonged immobilization can be used as animal
models of NTI.
Commentary. A large number of animal models have
been developed to study NTI (361–370). Essentially, the type
and extent of the bodily insult that triggers NTI in rats, mice,
and rabbits determines the magnitude of the alterations in
thyroid economy. These different animal models display
unique characteristics as to the timeline of the fall in serum T
4
and/or T
3
concentrations.
Different cell models have been developed to study the
modifications in deiodinase expression and thyroid hormone
action during NTI. Basically, different deiodinase-expressing
cell types are exposed to pro-inflammatory cytokines (e.g.,
interleukin-1b) and then evaluated for the expression of dif-
ferent components of thyroid hormone signaling such as TRb,
TRa, and deiodinase activity (361).
&
RECOMMENDATION 35b
The administration of bacterial LPS can be used to promote
central hypothyroidism, similar to that observed during
NTI, with reduction in hypothalamic TRH and a fall in
serum thyroid hormone levels.
Commentary. Intraperitoneal administration of LPS to
rats (371–373) or mice (173) has been used extensively to
promote central hypothyroidism and NTI (374). Despite the
drop in serum thyroid hormone levels, both TRH expression
in the paraventricular nucleus and serum TSH are decreased
in LPS-treated rats, resembling the euthyroid sick syndrome
(371,373). These LPS-induced changes are associated with an
elevation in tanycytes that may be key for the reduction in
TRH expression through feedback inhibition (373,374); this
effect is lost in the D2 KO mice (173). LPS batches vary in
efficiency; it is recommended to use a strain previously shown
to be effective; for example, O127:B8 Escherichia coli strain
in vivo in rats and mice (372–374). In vitro, treatment with LPS
increases D2 activity in human mesothelioma (MSTO-211H)
cells (375) and in cultured rat astrocytes (376).
[I] Assessing Thyroid Hormone Signaling at Tissue
and Cellular Levels
Overview. TRs mediate biological responses to thyroid
hormone by control of gene expression (377,378). TRs are nu-
clearreceptorsandbindspecicDNAresponseelementsin
genomic regulatory regions (enhancer elements) of target genes
(379–381). The DNA-bound TR can modify the activity of
chromatin remodeling complexes, RNA polymerase II, and the
basal transcriptional machinery to activate or suppress expres-
sion of a target gene. The transcriptional response is sensitive to
the concentration of ligand (T
3
)andtothedurationofexposure
to T
3
. This can be studied in animals or in cell models usually by
contrasting expression of a T
3
-responsive gene between hypo-
thyroid and thyrotoxic conditions (see Sections E and F).
The three canonical TR isoforms TRa1, TRb1, and TRb2
in mammals possess broadly similar transactivation
properties on many but not all response elements in vitro.
These isoforms mediate both overlapping and distinct
biological functions in mice in vivo (286,382,383); reflecting
cell-specific expression patterns, differences in isoform
expression levels, and possible isoform-specific structural
constraints in target gene recognition or cofactor interac-
tion (Table 3) (385–389).
The analysis of T
3
response genes is useful to address dif-
ferent types of research questions. T
3
response genes are
useful markers, or transcriptional endpoints, of the thyroid
hormone status of a tissue in animal models or in cell culture
lines in vitro. Analyses may be relatively simple, with a focus
on a few known target genes that reliably represent the tissue
status in response to T
3
. Exploratory large-scale (‘‘genome-
wide’’) screens of mRNA populations may be used to search
for new T
3
-response genes. At the molecular level, a specific
target gene can be investigated in depth to elucidate the
transcriptional mechanisms underlying T
3
action, entailing
the analysis of the DNA element that binds the TR and
functional analysis of these elements using transcriptional
assays, usually in heterologous cell culture systems. Ulti-
mately, the physiological relevance of such an element would
require in vivo evidence based, for example, on analysis of
transgenic reporter genes carrying the promoter and/or en-
hancer elements of the target gene.
Table 3. T
3
Receptor Genes and General Tissue
Expression Patterns of Receptor Isoforms
Gene
T
3
receptor
isoform Some main sites of expression
Thra
(Nr1a1)
TRa1 Pituitary, brain regions, heart,
intestine, bone, kidney, cartilage,
erythroid/lymphoid cells
Thrb
(Nr1a2)
TRb1 Pituitary, brain regions, heart, liver,
kidney, lung, cartilage, retina
TRb2 Pituitary, cartilage, cochlea, retina,
hypothalamus
INVESTIGATING THYROID HORMONE ECONOMY AND ACTION 119
[I.1] Gene expression as a marker of thyroid hormone status
Background. T
3
target genes can be either positively or
negatively regulated. Exploratory screens of cell lines or tis-
sues indicate that on average, approximately 50% of the T
3
-
responsive genes display increased mRNA levels and 50%
display decreased levels in response to T
3
(127,390–394). For
the majority of these genes, it remains unknown which re-
spond directly and which indirectly to T
3
, since in many tis-
sues relatively few genes have been defined as direct targets
by rigorous criteria that include the identification of the
functional TR binding sites in the gene. The mRNA levels
measured also represent the net outcome of other variables
such as the developmental stage of the tissue and the mixture
of cell types in the sample (395–397). Nonetheless, even when
these limitations are taken into account, the mRNA level of
selected genes provides a useful indicator of the response
status of the tissue. Independent methods such as Western
blot analysis are used to corroborate the data at the protein
level, and in situ hybridization analysis can define more pre-
cisely the cell types in which gene expression changes occur.
Any method used to confirm gene expression data requires
appropriately controlled analyses. For example, in Western
blot analysis, evidence for the specificity and sensitivity of the
antibody used is essential. Also, as a qualitative control for the
amount and integrity of protein sample loaded, the specific
protein band detected may be compared to an internal control
(reference) protein such as actin or RNA polymerase II, if
these proteins themselves do not vary substantially in re-
sponse to altered T
3
or TR status in the tissue being studied
(82,398). If necessary, the specific protein band detected may
also be quantified relative to the internal control protein band.
&
RECOMMENDATION 36
For general assessment of the tissue status in response to T
3
,
the analysis of representative known response genes pro-
vides informative data.
Commentary. The following genes are examples of T
3
response genes and serve as useful markers of tissue T
3
status:
(i) liver: Me1, Thrsp, Dio1, Gpd2, Cyp27a, and Fasn (390); (ii)
brain: Nrgn, Mbp, Hr, and Trh (paraventricular nucleus only)
(400,401); (iii) heart: Myh6, Myh7, and Atp2a2 (402); and (iv)
BAT: UCP1 (403). Determination of mRNA levels of selected
genes is accomplished by direct methods such as Northern
blot analysis (Fig. 12) or by indirect analyses based on am-
plification of cDNA using RT-qPCR if precautions are taken to
ascertain the specificity and quantitative validity of the PCR
assay (404). However, it should be noted that although these
genes can be useful, representative indicators of T
3
status of a
tissue, different genes in the same tissue may reflect a range of
direct or indirect mechanisms of response. Other factors in-
fluence the response of a given mRNA to T
3
in a specific tissue.
For example, Me1 is responsive to thyroidal status in liver, but
not the brain. Additionally, Me1 in liver is only regulated by
thyroid hormone after postnatal day 15 (405). Such temporal
and tissue specific considerations should be taken in to ac-
count when designing experiments that quantify mRNA.
[I.2] PCR analysis of mRNA expression levels
Background. PCR analysis of mRNA levels for T
3
-
responsive genes follows standard methods. PCR is extremely
sensitive and has the potential to generate ostensibly positive
signals from very few molecules of mRNA. Thus, it is im-
portant to perform negative control reactions lacking cDNA
templates. For the valid detection of mRNA transcribed from
intron-less genes, reverse transcription should be also per-
formed in the absence of the reverse transcriptase enzyme
(minus RT control) to exclude false-positive amplicons gen-
erated from a contaminating genomic DNA template. Given
these limitations, results should be supported independently
with more direct assays such as Northern blot, Western blot,
or in situ hybridization analyses.
&
RECOMMENDATION 37
A current method of choice to study the expression of T
3
-
responsive genes is RT-qPCR, which provides precise
quantification of mRNA levels, if appropriately validated.
Care should be taken to ensure that internal control genes
(i.e., reference genes) are not themselves T
3
responsive in
the system being studied.
Commentary. Care should be taken to ascertain that the
quality of the input RNA is adequate. Poor quality RNA can
limit reverse transcriptase efficiency and cDNA yields (406).
FIG. 12. Dio1 (type 1 deiodinase) is a representative T
3
-
responsive gene in the liver. Northern blot analysis showing
Dio1 mRNA levels in response to normal diet (ND), hypo-
thyroid (LID, low-iodine diet and antithyroid agents), and
hyperthyroid (0.5 T
3
) conditions. Dio1, upper panel, and
control glyceraldehyde-3-phosphate dehyrogenase (G3pdh),
lower panel. In WT mice (+/+), Dio1 mRNA is suppressed
by hypothyroid conditions and induced by hyperthyroid
conditions. Induction is defective in Thrb-deleted mice ( -/-).
Treatments: ND, normal diet; LID, hypothyroid groups
(0.05% methimazole [MMI] and 1% potassium perchlorate in
drinking water and low iodine chow, for 4 weeks); 0.5 T
3
,
hyperthyroid groups (same as LID but with T
3
added to
drinking water at concentration of 0.5 mg/mL for an addi-
tional 8 days or more). Quantitation: Dio1 mRNA level
in each condition is noted numerically below each lane rel-
ative to the level in +/+ mice on normal diet (assigned a
value = 1.0). Dio1 value is normalized to control G3pdh signal;
UD, undetectable Dio1. Signals were quantified by pho-
phorimager analysis of major bands.
120 BIANCO ET AL.
In general, standard primer design rules apply to RT-qPCR
except that amplicon size should be restricted to 50–200 bp
with a preferred amplicon size around 100 bp to maximize ef-
ficiency (407). When possible, primer pairs should span exon
exon junctions in cDNA to avoid amplification of potentially
contaminating genomic DNA. Primer design websites (e.g.,
Primer3, http://frodo.wi.mit.edu or www.ncbi.nlm.nih.gov/
tools/primer-blast) and databases of validated primer sets (e.g.,
http://pga.mgh.harvard.edu/primerbank) are freely available.
Two common detection methods in RT-qPCR use the SYBR
Green DNA binding dye or 5¢-nuclease (i.e., TaqMan probe)
assays (Fig. 13). In the unbound state, SYBR fluorescence is
negligible, but upon binding to double-stranded DNA its
fluorescence increases. The 5¢-nuclease assay takes advantage
of the intrinsic 5¢-exonuclease activity of many Taq polymerases
to degrade a fluorescence resonance energy transfer–linked
oligonucleotide probe designed to anneal within the amplicon.
Probe cleavage separates the 3¢-quencher fluorophore from the
5¢-reporter fluorophore. A real-time PCR machine measures
the accumulation of released fluorescent reporter product. The
5¢-nuclease assay offers possibly increased specificity over
SYBR Green detection as it avoids potential complications with
primer dimer formation (408). To validate the assay it is also
necessary to confirm the size of PCR band and/or the presence
of a single peak in the dissociation curve (Fig. 13C).
Quantification of mRNA by RT-qPCR usually requires
normalization to an internal control reference gene. Care
should be taken in selecting reference genes for thyroid re-
search because several reports have indicated that commonly
utilized reference genes (e.g., Actb, Gapdh) are somewhat
responsive to thyroid hormone in some specific tissues (409
412). The genes Rn18S and Ppia (cyclophillin A) appear suitable
for several tissues including brain, pituitary, and liver (351,413).
The abundance of a target mRNA is calculated by either rel-
ative or absolute quantification (406,414). In relative quantifica-
tion, the PCR critical threshold (Ct) of a test RNA sample (e.g.,
treated with T
3
) is compared to a reference sample (i.e., vehicle
control). Two methods of relative quantification are widely
used: DDCt-method and the relative standard curve method.
The DDCt-method of normalization makes use of a refer-
ence gene, such as a house keeping (Actb, Gapdh) or other
suitable forms of reference gene, that is expressed similarly
across all samples with minimal variation. The DDCt is cal-
culated as follows:
DCt
control
¼ Ct
target
Ct
reference
DCt
test
¼ Ct
target
Ct
reference
DDCt ¼ DCt
control
DCt
test
Relative Expression ¼ 2
DDCt
where Ct is the PCR cycle number at which the reporter
fluorescence is greater than the background signal (414).
In the relative standard curve method, cDNA from an in-
dependent sample is used to create a relative standard curve
utilizing the known mass of input RNA against which
FIG. 13. An example of using RT-qPCR to analyze a T
3
-responsive gene. The qPCR amplification plots indicate change in
the mRNA level for the T
3
-responsive gene OPN1LW/MW (red/green opsin) in the human retinoblastoma cell line WERI
following treatment with increasing T
3
concentrations spanning the physiological range. The increased SYBER Green de-
tection (DRn; SYBER Green fluorescence [reporter] normalized to background) following each PCR cycle demonstrates the
accumulation of a PCR amplicon. (A) Amplification plots of the internal control (reference) gene GAPDH (glyceraldehyde-3-
phosphate dehydrogenase). Note that at the critical threshold (Ct, dotted line across graph) no difference is detected between
the T
3
doses for GAPDH indicating that this reference gene is not responsive to T
3
treatment. (B) Amplification plots of the T
3
target gene OPN1LW/MW. Note that as the T
3
concentration increases, the PCR cycle needed to reach the threshold decreases
indicating the presence of higher levels of mRNA induced by T
3
. A plateau in T
3
-induced expression of OPN1LW/MW is
reached at 50 nM T
3
. (C) Melting curve analysis of OPN1LW/MW qPCR amplifications. Note that a single melt peak is
observed in a plot of the first negative derivative (-R; fluorescence over time; i.e., the change rate) against temperature
indicating that the increase in SYBER Green fluorescence detected is likely derived from a single PCR amplicon. Data are
unpublished observations from D.S. Sharlin and D. Forrest and are consistent with previous reports (752).
INVESTIGATING THYROID HORMONE ECONOMY AND ACTION 121
unknown samples, such as reference and test samples, can be
quantified. To control for potential variation in RT input
RNA, calculated masses can be normalized by the calculated
mass of an endogenous control (408).
In absolute quantification, the Ct for the mRNA of a given
gene is fitted to the Ct from a standard curve generated using
serial dilutions of samples that contain known quantities of the
target gene (i.e., plasmid DNA or in vitro transcribed RNA) (406).
The abundance of an mRNA may also be determined by the
so-called semiquantitative, or end-point RT-PCR, in which
amplicons are visualized by agarose gel electrophoresis after a
predefined, arbitrary number of PCR cycles (404). Although
arbitrary end-point RT-PCR can detect obvious changes in gene
expression, it suffers from low precision, and often involves
visualization of PCR products at the plateau phase of the
reaction, when the data do not accurately reflect the quantity of
initial starting mRNA material (407).
[I.3] Genome-wide analysis
of thyroid hormone–responsive mRNA
Background. Microarrays or genome-wide analyses by
other methods, such as next-generation sequencing of expressed
RNA populations (i.e., RNA sequences [RNAseq]), generate
large datasets that demand major computational investigation
to extract meaningful data for relevant genes. Typically, thou-
sands of mRNA sequences display some degree of change in
levels. Follow-up analysis by other methods is required to de-
termine if candidate genes are direct targets of T
3
action.
&
RECOMMENDATION 38a
Microarray or other methods of genome-wide screening for
mRNA expression patterns can be used for large-scale
analysis of gene expression changes in response to ma-
nipulations of thyroid hormone signaling pathways. Can-
didate transcripts of interest must be corroborated by
independent methods, such as RT-qPCR, Western blot, or
in situ hybridization analysis.
Commentary. Many studies have been published featur-
ing microarray analysis of a given tissue or cell line following
T
3
treatment of an intact organism or cultured cells. Specific
techniques for genome-wide expression analysis are evolving
rapidly, and multiple microarray systems and data mining
software packages exist and can be used as desired. Most
post-array analytical software platforms allow for manipula-
tion of the stringency parameters that set a threshold for sig-
nificantly changed genes, such as fold-change, false discovery
rate (to correct for multiple comparisons), and p-value. In
general, at least three biological replicates per treatment
group (e.g., T
3
exposure) and control group (e.g., vehicle) are
required for ‘statistically significant’ differences to be de-
tected if statistical significance is taken to mean a p value of
< 0.05, false discovery rate of <0.05, and fold-change >2 (417).
Given the high likelihood of false-positive differences, results
of microarray studies should best be viewed as ‘hypothesis-
generating data’ in nature, and changes in key transcripts
must be confirmed via PCR and other methods.
&
RECOMMENDATION 38b
RNAseq uses deep-sequencing technology and can pro-
vide a more precise measurement of mRNA and variant
mRNA isoform levels compared to microarrays as data
derived from RNAseq is not biased by analysis using pre-
determined genes and probes.
Commentary. RNAseq requires major bioinformatics
analysis of sequence reads (418). The inclusion of extrane-
ous cell types in a tissue sample should be minimized, as
qualitative variations between samples will increase experi-
mental ‘noise’ and reduce the resolution of analysis. Tissue
sampling may be refined using laser microdissection to isolate
defined pieces of tissue from histological sections or fluores-
cence-activated cell sorting to isolate enriched cell popula-
tions based on specific cell markers (419,420). Observed
changes in mRNA expression patterns might represent indi-
rect alteration in mixed cell populations in tissues (421,422). A
cell line in culture, although not a physiological sample, may
provide a more homogeneous cell population for certain an-
alyses.
[I.4] Mechanisms of gene regulation by thyroid hormone
Background. The TR binds to DNA elements known as T
3
responsive elements (T
3
REs) (423). In general, TR binds the
T
3
RE as a homodimer or heterodimer with retinoid X recep-
tors (RXRs). These receptor complexes attract co-repressors or
co-activators depending on whether the TR is in a T
3
-bound or
unbound state (424–426). Analyses of TR–DNA interactions
often use two broad approaches: (i) transcription response
(‘‘transactivation’’) assays, typically in transfected cells in
culture, and (ii) assays to investigate TR binding to DNA on
isolated DNA response elements (electrophoretic mobility
shift assay [EMSA]) or on genomic DNA in a chromatin state
(chromatin immunoprecipitation [ChIP]).
&
RECOMMENDATION 39
The regulation of a target gene by TR can be investigated in
tissue culture cells co-transfected with a reporter gene
plasmid plus a TR-expressing plasmid. Transactivation of
the reporter is assessed in response to added T
3
at varying
doses over a physiological range and can be tested in dif-
ferent cell lines since responses may be influenced by host
cell-specific factors.
Commentary. The candidate T
3
RE or enhancer region of
the target gene with its natural promoter or an artificial pro-
moter is ligated to a readily detectable reporter gene, usually
firefly luciferase. Luciferase activity generates light, which is
detected using a luminometer (427). Promoter-less and
enhancer-less luciferase reporters are used as controls to ex-
clude the possibility of spurious responses originating from
‘stealth sequences’ fortuitously present in the luciferase gene
or vector backbone (428).
Transfection efficiency is assessed by co-transfection of an
internal control reporter plasmid usually with Renilla lucif-
erase driven by a constitutively active promoter. The specific
readout of a reporter assay is normalized as firefly luciferase
activity (test gene) over Renilla luciferase activity (internal
control gene). It is imperative to establish that internal control
reporters do not respond to the experimental conditions be-
cause this can inappropriately distort the readout ratio (429).
Consideration should be given to whether the reporter
vector is chromatin-forming or non–chromatin-forming.
122 BIANCO ET AL.
Standard or ‘naked DNA’ reporter plasmids such as pGL4
often yield useful data in transiently transfected cells. Alter-
natively, vectors that carry an episomal replication origin,
such as pREP4, can form chromatin in transiently transfected
cells. Some DNA response elements require a chromatin-
forming vector to display activity (430).
The location of the promoter for a given target gene of
interest should be verified. Not all genes represented in the
genome databases depict an accurate 5¢ gene structure, which
requires experimental mapping of 5¢ ends of mRNAs from
different tissues. The most upstream exons of a gene (trans-
lated or untranslated) can be identified by RNAse protection
assays or 5¢-rapid amplification of cDNA ends analysis
(431,432). The function of a presumptive promoter should be
demonstrated using luciferase assays in transfected cells or
reporter transgenes in an animal model. The preconception
should be avoided that relevant enhancers reside only up-
stream of the promoter of a target gene. Many enhancers re-
side within introns or downstream of the gene (433).
&
RECOMMENDATION 40a
EMSA can demonstrate direct binding of the TR to a DNA
element thereby providing evidence that a gene is a direct
target of TR. EMSA is based on the observation that
protein-bound DNA migrates at a slower rate than un-
bound DNA when subjected to electrophoresis. The DNA
element (the probe) is usually radiolabeled to allow sensi-
tive detection of protein:DNA probe complexes.
Commentary. TR binding sites are typically related to the
consensus motif for a nuclear receptor binding site, 5¢-
AGGTCA-3¢ and often occur in dimeric repeat configurations
(423,434). Optimal EMSA probes are short double-stranded
oligonucleotides (*20–30 bp). Longer probes (or degraded
probes) may increase nonspecific binding, which appears as
indistinct, smeared signals after EMSA. Radioactively end-
labeled probes offer great sensitivity and allow detection of
protein–DNA complexes using x-ray film or phosphorima-
ging. Alternatively, probes can be labeled nonradioactively
(e.g., with biotin), followed by secondary detection with
streptavidin and enzymatic substrates similar to those used
for Western blotting (435).
Nuclear or whole cell extracts from the tissue or cell line of
interest provide a source of TR protein. Lysates of cultured
cells transfected with a TR-expressing plasmid are another
common source of TR protein (381). Alternatively, TR protein
can be generated using a cell-free in vitro translation system
(436,437).
The specificity of the TR–DNA probe interaction is con-
firmed by two standard control tests: (i) the use of an anti-TR
antibody that supershifts or disrupts the retarded band (a
parallel negative control using a nonspecific antibody should
not disrupt the specific TR–DNA probe complex), and (ii) the
specificity of the TR binding to the labeled DNA probe should
be confirmed by competition in the presence of an excess of
unlabeled ‘cold’ oligonucleotide probe. Two types of com-
petitor oligonucleotide probe are used in parallel samples:
wild-type probe and probe containing point mutations within
the proposed TR binding site. A dose-dependent reduction in
the intensity of the shifted band signal is expected when using
cold wild-type probe but not mutant probe (436). Typically,
each cold competitor is added in two or three doses, re-
presenting a range of approximately 2- to 100-fold excess over
the labeled probe.
&
RECOMMENDATION 40b
ChIP can be used to indicate that the TR associates with a
target gene in its natural context in the genome in a tissue or
cell line. The most critical requirement for ChIP is a high-
quality specific antibody for immunoprecipitation, with the
specificity being established by control experiments.
Commentary. ChIP is used for testing binding of tran-
scription factors to known target genes or for exploration of
previously undefined binding sites. A positive ChIP result
indicates an association with a region of chromatin DNA but
does not differentiate between direct binding to DNA and
indirect binding to a complex of other factors that bind DNA.
Currently, there is little evidence that available antibodies
against TR generate consistently reproducible ChIP results for
mammalian tissues, although some ChIP data have been re-
ported for cell lines and amphibian tissues (438–442). The
difficulties in generating ChIP data on natural tissues may be
a consequence of relatively low concentration of TR in
mammalian tissues compared to other types of more abun-
dantly expressed transcription factors.
Optimal conditions for cross-linking and sonicating chro-
matin should be established empirically for any tissue or cell
line (443). A negative control for ChIP utilizes purified non-
specific IgG, pre-immune serum, or antibodies against foreign
proteins not present in the tissue. Genomic DNA fragments
isolated by ChIP are identified by PCR using a series of primer
sets that span the genomic region of interest as well as nega-
tive regions of the gene or distant genomic regions (444). All
negative controls can generate a background signal and
should be tested empirically.
&
RECOMMENDATION 40c
More advanced applications of ChIP, demanding stringent
technique and computational analysis, involve the combi-
nation of ChIP with microarray (ChIP-chip) or next-
generation sequencing to identify genome-wide binding
sites for a nuclear factor.
[I.5] Mouse models for indicating thyroid hormone
and TR signaling in tissues
Background. Numerous KO and knockin mouse strains
derived by genetic manipulation have yielded a wealth of
information on the physiological functions of TR isoforms
in vivo. Genetic manipulation has also been used to generate
models that yield insights into where and when T
3
or specific
TR proteins are present in tissues in vivo. There is scope for
further development of such approaches, to elucidate at the
cellular level, the basis of T
3
actions in vivo.
It should be borne in mind that the lack of receptor in KO
mice may produce certain differences in phenotype compared
to a lack of hormone. For example, mice lacking all T
3
receptors
(TRa1, TRb1, TRb2) display multiple tissue phenotypes but are
not as small or as retarded as mice with severe, congenital hy-
pothyroidism (75). There may be several explanations for these
differences. An explanation that has been supported by study of
the cerebellum in vivo, is that in hypothyroidism, the TR exerts
chronic ligand-independent dysregulation of gene expression
INVESTIGATING THYROID HORMONE ECONOMY AND ACTION 123
to produce a more severe outcome than the absence of a TR. In
the absence of TR, this ligand-independent dysregulation of
gene expression could not occur such that the phenotype is less
severe (445). However, few tissues and few target genes have
been studied in detail to indicate how widely this explanation
may apply. In summary, the deletion of TR isoforms yields
precise information on the tissue-specific functions of TR iso-
forms. However, the phenotypes may not always be reflected in
hypothyroid models. Conversely, hypothyroid phenotypes
maynotbefullyreectedinTR-decientmodels.
&
RECOMMENDATION 41a
Mouse strains with targeted KO or knockin mutations in
the endogenous Thra and Thrb genes provide models to
study TR functions in vivo.
Commentary. Numerous mutations in the Thra and Thrb
genes (also known as Nr1a1 and Nr1a2, respectively) have
been derived but only a brief mention of phenotypes is given
here. For more detailed reviews on Thra and Thrb mutations,
readers are referred elsewhere (286,382,383) and to the Mouse
Genome Informatics website (www.informatics.jax.org). An-
other useful online resource is the Nuclear Receptor Signaling
Atlas (www.nursa.org).
The initial gene targeting studies established that Thrb is pri-
marily responsible for the regulation of the HPT axis (446,447),
the development of the auditory and color visual systems
(448,449), and the majority of T
3
actions in liver (390,393,450).
The Thra gene is primarily responsible for determining thermo-
genic and cardiac functions (386,451) and the maturation of the
intestine (452), bone (453), and certain brain tissues (445,454).
In contrast to KO models, knockin mutations in the coding
sequence of the Thra or Thrb genes have generated mouse
models that express dominant negative TRa1 and TRb pro-
teins with little or no response to T
3
(456). These mutations
create models of localized ‘tissue-restricted’ hypothyroidism
selective for tissues in which TRa1orTRb is the predominant
T
3
receptor isoform expressed.
&
RECOMMENDATION 41b
Mice devoid of known T
3
receptors (TRa1, TRb1, and TRb2)
can be used to model the complete absence of TR-mediated
signaling.
Commentary. Mice lacking specifically TRa1, TRb1, and
TRb are viable but have cellular and functional defects in mul-
tiple organ systems, indicating that in many tissues, TRa1and
TRb isoforms serve additional, common functions that were not
evident in models with single receptor gene mutations (75,457).
&
RECOMMENDATION 41c
Coding sequence changes introduced into both Thrb and
Thra genes have provided mouse models for the human
syndromes of resistance to thyroid hormone.
Commentary. The human syndrome of resistance to
thyroid hormone is typically caused by heterozygous coding
mutations in the THRB gene that generate dominant negative
proteins with little or no response to T
3
(458). Recently, similar
mutations in the human THRA gene have been identified that
generate dominant negative TRa1 proteins (459,460). The Thrb
and Thra knockin mouse models display a range of tissue-
selective phenotypes and offer the opportunity to investigate
the cellular and molecular defects underlying the disease
symptoms in specific tissues (456,461–465).
&
RECOMMENDATION 41d
Transgenic reporter mice that express a T
3
-responsive chi-
meric protein provide an in vivo model that allows for
monitoring of the presence of T
3
in various tissues.
Commentary. It has traditionally been difficult to detect T
3
at the cellular level in natural tissues, especially in specialized
cell populations. RIAs performed on tissue homogenates
provide no information on which cell types contain T
3
.The
transgenic FINDT
3
mouse model offers the advantage of
detection of T
3
in localized regions within complex tissues
(278,467). These mice express a chimeric protein consisting of a
yeast Gal4 DNA binding domain fused to a TRa1 ligand-
binding domain. The expression of the reporter does not
require endogenous TR. The presence of T
3
is detected visually
on tissue sections based upon activation by the chimeric pro-
tein of a Gal4-responsive promoter-fused to a LacZ reporter
gene (Fig. 14).
This approach can be used to detect T
3
activity in several
tissues, but reportedly has limited sensitivity in certain brain
regions (278). Using beta-gal as the readout in transgenic
mice, T
3
signaling was absent in the early embryo and was
then detected at around E11.5–E12.5 in different primordia
(i.e., CNS, intestine, etc.). Since at this time, fetal thyroid
function in the mouse is still inactive, and these early signals
may reflect maternal T
4
or T
3
activity. Early T
3
signaling was
observed in the brain (i.e., diencephalic primordia, medulla
oblongata) and sense organs primordia (otic vescicle, olfac-
tory epithelium, retina), whereas at late stages (E15.5–E17.5)
beta-gal expression was localized in other primordia (i.e., in
the bones, follicular nerves of the vibrissae, as well as in the
small intestine primordia, but also in the medulla oblongata)
(468).
&
RECOMMENDATION 41e
Targeted insertions in the endogenous Thra and Thrb genes
generate TR proteins fused to protein tags that can be used
to monitor expression of TR isoforms in specific cell pop-
ulations in vivo.
Commentary. The fused tag allows detection of TR iso-
form expression by immunohistochemistry (IHC) or immu-
nofluorescence in tissue sections or whole-mount preparations
and offers the possibility of detection at single cell resolution.
Thra1
gfp
mice carry green fluorescent protein (GFP) fused
in-frame at the C-terminus of the intact TRa1protein(469).
Detection of TRa1-GFP protein has been reported in specific
brain tissues and cell types. LacZ inserted into the Thrb gene
allows detection of TRb isoforms (430,449,457) and has re-
vealed TRb2 expression in restricted cell populations in the
cochlea, pituitary gland, and cone photoreceptors.
[J] Assessing Thyroid Hormone Signaling
by Way of Systemic Biological Parameters
Overview. Thyroid hormones exert diverse actions in
virtually all tissues during development, infancy, adoles-
cence, and adulthood in areas such as growth, cognition,
124 BIANCO ET AL.
skeletal-muscle homeostasis, cardio-circulatory function, en-
ergy homeostasis, and intermediary metabolism. It is essential
to adopt a physiological whole animal approach for investi-
gation of thyroid hormone action in vivo. By necessity, this
requires longitudinal studies of T
3
-responsive biological pa-
rameters and/or biochemical markers.
Studies generally focus on the trans ition from hypothy-
roidism to euthyroidism and thyrotoxicosis. Two comple-
mentary strategies can be adopted: (i) collect sequential
biological measurements from the same animals at baseline
and intervals during the period of study, or ( ii) run control
and experimental groups in parallel and collect biological
measurements at specific time points; this model has
the advantage that animals can be killed at different time
points and tissues harvested for structural or biochemical
analyses.
In general these studies are performed in common rat or
mouse strains. More advanced animal models include geneti-
cally modified mice in which thyroid hormone signaling has
been disrupted or modified by either global or tissue-specific
gene targeting. With such mouse models it is important to
ensure experiments are designed to avoid the confounding
influence of mixed genetic background if possible. Ideally,
mutant strains should be backcrossed to homogeneity onto a
suitable background such as C57BL/6. If this is unrealistic (e.g.,
if crossing CRE and FLOXED lines from different genetic
backgrounds), littermates should be compared and adequate
power calculations performed. Males and females may need to
be studied separately because responses to thyroid hormones
can be sexually dimorphic.
T
3
-dependent biological parameters can also be studied
using ex vivo preparations, such as freshly isolated cells, cul-
tures of primary cells or established cell lines. In general, such
in vitro systems have poor responsiveness to T
3
, with only a
few being well characterized. However, cell-based models
have been invaluable in the deconstruction and understand-
ing of molecular mechanisms underlying the complex bio-
logical effects of T
3
.
[J.1] Central nervous system
Background. Most effects of thyroid hormone in the brain
are observed during development. Endemic neurological
cretinism results from maternal iodine deficiency and the
consequent maternal hypothyroxinemia, in which circulating
T
4
levels are low for the stage of pregnancy. Maternal hy-
pothyroxinemia causes neurological thyroid hormone defi-
ciency in the developing fetus resulting in mental retardation,
spastic diplegia, deaf-mutism, and squint in the absence of
general signs of hypothyroidism. Even though endemic cre-
tinism can be mitigated by iodine supplementation to prevent
or correct first trimester maternal hypothyroxinemia, iodine
deficiency remains the commonest cause of preventable
mental retardation. On the other hand, neurological features
of neonatal hypothyroidism may be less profound because
they are dependent on the severity of hypothyroidism. Ab-
normalities are largely preventable by immediate thyroid
hormone replacement, although deficits in memory and IQ
may persist. Nevertheless, untreated neonates have growth
retardation and general features of hypothyroidism with
mental retardation, tremor, spasticity, and speech and lan-
guage defects. The differences between endemic cretinism
and congenital hypothyroidism illustrate that the timing of
thyroid hormone action is fundamental for neurodevelop-
ment (470,471).
Developmental hypothyroidism also causes sensory im-
pairment, including permanent deafness, which in rodent
models has been shown to involve deformity in the inner ear
tectorial membrane and impaired maturation of the sensory
epithelium (247). There is also a visual phenotype since thy-
roid hormone affects the development of color vision. In
adults, thyroid hormone acts on mood, behavior, and cogni-
tive function, and sophisticated tests have been validated to
investigate the effects of thyroid hormones on behavior phe-
notypes in mice. Thyroid hormone also affects muscle
strength and fatigue as well as the motor and cerebellar sys-
tems. Individuals with hypothyroidism are sluggish and fa-
tigued with muscle weakness and slow-relaxing reflexes,
whereas in thyrotoxicosis there is tremor, muscle wasting,
hyper-reflexia, and fatigue.
FIG. 14. T
3
signaling represented by b-galactosidase stain-
ing (blue) in coronal brain sections in postnatal day 5 mice
carrying the FINDT
3
reporter transgene. (A, D) FINDT
3
reporter on wild-type (WT/FINDT3) mice; (B, E) FINDT
3
reporter in mice lacking type 3 deiodinase, a thyroid hor-
mone-inactivating enzyme (D3KO/FINDT3). (C, F) No b-
galactosidase staining was detected in sections from control
mice (WT) not carrying the transgene. S, septum; Pi, piriform
cortex; MCx, motor cortex; SCx, sensory cortex. Note in-
creased b-galactosidase activity in D3KO/FINDT
3
mice in
the piriform, motor, and sensory cortex, consistent with in-
creased T
3
exposure in mice lacking type 3 deiodinase.
Cryosections (50 lm thick) were stained using 1 mg/mL of
X-gal (5-bromo-4-chloro-3-indolyl-d-galactopyranoside), a
colorigenic substrate for b-galactosidase. Reproduced and
adapted from Hernandez et al. (278) with permission. ª 2013,
The Endocrine Society.
INVESTIGATING THYROID HORMONE ECONOMY AND ACTION 125
The adult brain is also responsive to thyroid hormone. A
standardized observational screening assessment of behavior
and function in rodents has been described (SHIRPA) that
allows a general and comprehensive neurological (muscle
and lower motor neuron, spinocerebellar, sensory, and auto-
nomic function) and neuropsychiatric (activity, learning,
arousal, fear, aggression, feeding, irritability, etc.) analysis to
be performed prior to more detailed investigation of abnor-
mal parameters and neurological pathways (472). A modified
and standardized SHIRPA assessment has been adopted by
the International Mouse Phenotyping Consortium as part of
the initial characterization primary phenotype screen of
mouse mutants generated by the International Knockout
Mouse Consortium (473,474). Such an approach is being
adopted to include phenotype analysis of all physiological
systems and facilitate standardization and eventual genera-
tion of comprehensive and large repositories of phenotype
data.
&
RECOMMENDATION 42
Rodent models can be used to study how thyroid hormone
affects brain development and function. The correlation
between the time of onset of fetal thyroid function and the
chronologic age is different in rodents than in humans;
the stage of neurobiological development rather than the
chronological age of the animal should be considered.
Commentary. Congenital hypothyroidism promotes
profound impairment of brain development and function. For
example, selective and persistent cognitive problems may be
seen in children that were born with congenital hypothy-
roidism. Sophisticated imaging, such as magnetic resonance
imaging (MRI), has shown that children and adolescents with
congenital hypothyroidism have reduced hippocampal size
and abnormal hippocampal growth patterns relative to peers.
More importantly, reduced hippocampal volumes predict
poor memory performance (475).
Rodent models of congenital hypothyroidism have been
studied extensively, given that most anatomical and func-
tional components of the mammalian brain are similar be-
tween man and other mammalian species (476). However, in
many respects they do not constitute an ideal animal model
for study of fetal maternal thyroid economy during gestation.
The main differences involve the onset of fetal thyroid func-
tion and timing of neurodevelopmental events in relation to
birth. The issue is how much different species vary in relation
to man according to the proportion of the brain growth spurt
that is postnatal (476). In this respect, thyroid hormones do
not influence very early events such as neural induction and
the establishment of polarity during brain development. In-
stead, they modulate later processes, including neurogenesis
and dendrite proliferation, myelination, and synapse forma-
tion. The timing of onset of thyroid hormone action in the
developing brain is thus crucial and physiologically impor-
tant. Alternatively, sheep have been extensively used as an
animal model. The guinea pig is of potential interest as well
because its in utero thyroid maturation is much closer to that of
the human (477).
Many rodent studies have focused on the rat cerebellum
because its neurobiological development is predominantly
postnatal and thus more easily exposed to experimental ma-
nipulation (476). However, other regions including the cere-
bral cortex, basal ganglia, cerebellum, and hippocampus have
also been studied using standard histological approaches.
These techniques have also been adapted to study the spinal
cord and dorsal and ventral nerve roots (478). For example,
the Purkinje cell neurons and their spatial organization are
central to the function of the cerebellum. They are formed just
before birth in the rat and their number is not affect by hy-
pothyroidism at birth but their migration and maturation is
severely impaired (476). The effect of thyroid hormone defi-
ciency on maturation of the rat cerebellar cortex may still
be observed if hypothyroidism is induced up to the second
postnatal week (476). Similarly, most rat forebrain neurogenesis
is completed at birth. However, the structural analysis indicates
that there is marked hypoplasi a of the neuropil (476). Thus, it
seems clear that thyroid hormone affects direct brain maturation
through specific effects on cell differentiation. Thyroid hormone
slows down cell division while concomitantly stimulating the
onsetofcelldifferentiation.
&
RECOMMENDATION 43
Given the functional diversity and structural complexity of
the CNS studies of T
3
-responsive genes in the brain are
greatly enhanced if studied via in situ hybridization or IHC.
Commentary. Microarray analysis, Northern blotting,
RT-qPCR, next-generation RNA sequencing, and in situ hy-
bridization have all been used to study T
3
-regulated gene
expression in the brain (392,401,479). Protocols for in situ
hybridization and IHC have been described (445,480–484).
These may be performed on paraffin-embedded sections or
fixed frozen tissue if preservation of detailed histological
architecture is of significant importance (485). However,
paraffin embedding interferes with RNA quality, thus better
signal-to-noise ratio and analysis of gene expression are
obtained using fixed frozen tissue. For example, in situ
hybridization has been used to define TRH as a negatively T
3
-
regulated gene in the hypothalamus (486–489) and to study
the cortex/dentate gyrus expression of RC3/neurogranin,
which is positively regulated by T
3
(396).
When using quantitative methods, careful sample dissec-
tion is needed given the highly compartmentalized nature of
the brain. In some cases, pooled samples from 5–10 mice are
required to obtain sufficient RNA for such studies (485). A
more appropriate way of obtaining brain samples is through
laser-capture microdissection, which allows accurate isolation
of specific cell-types embedded in a heterogeneous tissue mi-
croenvironment. Laser-capture microdissection works under
micromanipulator-assisted direct microscopic visualization
and the samples obtained can then be processed for RNA
isolation (491).
A microarray analysis of hypothyroid mouse cerebral cor-
tex identified 316 genes positively regulated and 318 genes
negatively regulated by T
3
(127). The responsiveness of sub-
sets of these genes to T
3
was confirmed by RT-qPCR in a brain-
specific, severely hypothyroid mouse model, the double
Mct8/D2 KO mouse (127) and in the brain of systemically
thyrotoxic mice or D3 KO mice (492). Through the study of
systemic hypothyroidism or thyrotoxicosis, T
3
-regulated
genes have also been identified in specific regions of the brain
(e.g., cerebellum, cortex, hippocampus). In the developing
cerebellum, systemic hypothyroidism resulted in altered ex-
pression of 2940 genes, of which 1357 were up-regulated and
126 BIANCO ET AL.
1583 down-regulated as assessed by microarray analysis
(493). However, the number of cerebellar genes directly reg-
ulated by T
3
is likely to be much smaller as evidenced in
primary cultures of cerebellar neuronal cells studied via mi-
croarray RNA hybridization (494). In any such screen, strin-
gent statistical criteria should be used to narrow down the
number of candidate T
3
-regulated genes for further confir-
matory analyses.
Brain protein expression is typically analyzed by Western
blotting or IHC of fixative-perfused brains (post-fixed over-
night at 4C and cryoprotected in 30% sucrose). IHC is suit-
able for quantitative approaches only if the detection system is
kept unsaturated; quantification of secreted proteins and
peptides by this method is not recommended. Studies can be
performed on free-floating cryostat sections of specific brain
regions (478). The approach was used to study the effects
of congenital hypothyroidism on microtubule-associated
protein-2 expression in the cerebellum of the rat (495) and the
expression of thyroid hormone transporters in the cochlea
(496). In addition, IHC was also used to study the regulation
of reelin and dab1 by thyroid hormone in different brain re-
gions (497), whereas Western blotting and in situ hybridiza-
tion were used to study tenascin-c (498). These techniques
have differing advantages and disadvantages. Thus, immu-
nocytochemistry provides the ability to identify precisely
where in the CNS the antigen is located in individual cells or
at the subcellular level in the nucleus or cytoplasm or even
more precisely if electron microscopy is used. It also provides
an excellent method to verify loss or re-expression of a par-
ticular antigen in transgenic mouse models or following an-
atomical manipulation such as electrolytic lesioning or
transection. Because the content of an antigen in any partic-
ular cell can be influenced by its rate of synthesis, transport,
degradation, and secretion, in situ hybridization histochem-
istry tends to be a much better quantitative approach than
IHC. Alternatively, specialist microdialysis techniques can be
used to measure secreted protein concentrations. Western
analysis also can be subject to some of the same concerns with
respect to quantitation as immunocytochemistry, but has the
advantage of being able to identify the size of the antigen(s)
being identified.
&
RECOMMENDATION 44a
Functional hearing deficits resulting from thyroid hormone
deficiency may be studied by measuring auditory evoked
brainstem responses to determine stimulus intensity
thresholds.
Commentary. Analysis of middle ear anatomy involves
histological inspection of the ossicle for deformities (499).
Assessment of ossification requires staining with alcian blue
and van Gieson to visualize cartilage and bone formation.
Cochlear anatomy is examined on inner ear sections. Metha-
crylate plastic-embedded samples of paraformaldehyde/
glutaraldehyde-fixed cochlear tissue yield good preservation
of cellular structure; 3- to 4-lm-thick sections can be stained
with toluidine blue (448,501) or thionin (499,502) or aqueous
hematoxylin (277).
A relatively rapid, noninvasive means of assessing auditory
function is the measurement of the auditory-evoked brainstem
response (277,448,501,502). Investigation of other features of
cochlear function or cochlear nerve function may include de-
termination of the endocochlear potential, compound action
potential, and other physiological parameters (277).
&
RECOMMENDATION 44b
Deficits of visual function that result from thyroid hormone
deficiency may be studied by measuring visual evoked
responses in electroretinograms to determine wavelength
sensitivity and light intensity thresholds.
Commentary. Histological analysis of the retina is per-
formed on 2 lm sections obtained from methacrylate plastic-
embedded samples. The dorso-ventral axis of the globe is
marked. In situ hybridization and immunohistochemistry are
used to assess expression of candidate target genes and are
typically performed on 10–16 lm paraformaldehyde-fixed
cryosections (279,398,503,504). Protein expression analyses
can be performed on protein extracts prepared from dissected
retina to investigate various opsin and rhodopsin proteins
using Western blot and immunohistochemical studies
(279,398,503,504,506).
Cone and rod photoreceptor functions can be assessed by
analysis of the electroretinogram, which is typically recorded
in young adult mice. Given the critical functions of thyroid
hormone in different cone types, it is important to analyze
photopic (light-adapted) cone responses in response to spe-
cific light wavelengths that are selective for the different cone
populations, namely, those with peak sensitivity to medium-
long (M, or ‘green’ *520 nm) and short (S, or ‘blue’
*367 nm) wavelengths of light. Scotopic (dark-adapted) rod
responses should also be determined (279,398).
&
RECOMMENDATION 45
Thyroid hormone effects on neuron ion channels and nerve
conduction can be investigated by patch-clamp analysis or
microelectrode recording.
Commentary. Detailed analysis of neuronal function re-
quires specialist techniques best performed in collaboration
with established neuroscience laboratories. For example,
patch-clamp analysis of whole neurons in tissue slices or
microelectrode recording from tissue slices have been used
to study the effects of thyroid hormones on neuron function
(478). Whole-cell patch-clamp recordings of individual neurons
can be performed using specialist instruments on 300 lmpara-
sagittal brain slices obtained from mice. Neurons are visual-
ized by infrared-differential interference contrast microscopy
to allow selection of cells to be recorded (478). Advanced re-
cording procedures can measure hypothyroidism-induced
changes in long-term potentiation in hippocampal neuron
populations, a cellular mechanism of synaptic plasticity that is
thought to be involved in memory (507–510).
&
RECOMMENDATION 46a
Standardized neurological tests can be used to investigate
how thyroid hormone affects neuromuscular control at
both motor and cerebellar levels. The use of complemen-
tary methods is recommended.
Commentary. Open-field testing protocols have been
used to quantitate locomotor activity for distance travelled
and speed over a defined period (187,483,511) in order to in-
vestigate both behavioral and stamina aspects of
INVESTIGATING THYROID HORMONE ECONOMY AND ACTION 127
neuromuscular function in response to alterations in thyroid
status. Motor and muscle function can be investigated by
standardized tests of strength, including grip strength deter-
mined using an automated grip strength meter, an electronic
pull strain gauge, or a hanging wire test (478). Accelerating
rotarod testing of balance and motor coordination is per-
formed by determining the time an individual mouse is able
to stay on the rod during its rotation following a defined
training period prior to testing (187,483). Beam walk testing
using a series of elevated beams of differing widths may be
used to determine balance by measurement of foot slippage
and time to cross the length of the beam (478). Alternatively,
vertical pole tests are used to investigate agility and balance
by determining the time taken for a mouse to invert and run to
the base of a pole when placed near the top (187). Footprint
analysis of gait can be assessed in mice trained to walk along a
filter paper with their feet painted with different color non-
toxic paints applied to forelimbs and hindlimbs. Determina-
tion of stride length, front and hind base width, interstep
distance, heel usage, and hind paw angle can be used to assess
balance and coordination (478).
&
RECOMMENDATION 46b
Mood, behavior, learning, and memory are responsive to
thyroid hormone and can be investigated using established
behavioral tests. The use of complementary methods is
recommended.
Commentary. Features of depression are investigated by
learned helplessness testing. In this situation, an operant
learning system, comprising a conditioning chamber with a
shock generator, allows analysis of ‘active avoidance’ be-
havior by discriminating avoidance from escape (511). In
order to investigate anxiety a ‘passive avoidance’ system
should be used. A light/dark box is employed to determine
the time spent in light versus dark during a defined total time
period (511–513). An elevated plus sign–shaped maze, con-
sisting of two opposed open arms and two opposed enclosed
arms, can be used as an alternative assessment. In this situa-
tion, the time spent in exploratory behavior in the open arms
versus nonexploratory behavior time in enclosed arms is de-
termined (187,483). Startle responses as further indicators of
anxiety and restlessness may also be investigated (511).
To investigate the effects of altered thyroid status on
learning and to measure effects on recognition memory,
studies in an open field with familiar and novel objects can be
employed in novel object recognition tasks (187,483). Open-
field testing can also be used to investigate features of de-
pression because it evaluates the tendency of mice to explore
openly or stay frozen at the edges of the field or to present
rearing behavior. Additional tests to investigate learning and
memory involve the use of a Morris water maze filled with
opaque water and containing a small platform located in one
quadrant just below the water surface. The mouse is placed on
the small platform for 10 seconds and then placed in the
water. The time taken to find and climb onto the platform is
recorded. The task is repeated on sequential days to investi-
gate learning and memory. The visible cue test replicates the
experiment except that the platform is clearly visible (187).
The water escape test investigates visual awareness by plac-
ing a visible escape ladder on the side of the vessel and re-
cording the time taken to climb on to the ladder (187).
&
RECOMMENDATION 47
The hypothalamic effects of thyroid hormone on behav-
ioral and metabolic parameters can be studied directly after
intra-cerebro-ventricular (ICV) administration of T
3
.
Commentary. Thyroid hormone exhibits a number of ef-
fects that are mediated directly at the medial basal hypo-
thalamus. An example is the increase in food intake that
rapidly follows administration of T
3
directly into the ventro-
medial nucleus (514–516). A related approach is to implant T
3
crystals stereotaxically into discrete hypothalamic areas. This
method has been used to define the direct role played by T
3
in
the TRH negative feedback mechanism (487,488) and to study
the photoperiod-mediated regulation of seasonal energy bal-
ance and reproduction in the Siberian hamster (517). Fur-
thermore, the direct role played by a number of neuropeptides
on the HPT axis has been studied using the ICV route; for
example, aMHS, neuropeptide-Y, and agouti-related protein
(518–522).
&
RECOMMENDATION 48a
Brain cell models that respond to thyroid hormone in vitro
can be used to model T
3
effects in the CNS.
Commentary. Primary cultures of hippocampal neurons
and neuronal cell lines respond to T
3
in vitro. In cultured
neurons, exposure to T
3
significantly increases the neurite size
and length, as well as acetylcholinesterase activity (523). In
addition, T
3
directly affects the development of cultured cer-
ebellar Purkinje cell dendritic processes through activation of
TRa1 (524). T
3
can also enhance neuronal differentiation in-
duced by retinoic acid treatment of embryonic stem cells
(525). Glial cells are also responsive to T
3
(526). For example,
the activity of the cell maturation marker glutamine synthe-
tase increases in cultured cerebellar astrocytes in response to
T
3
(527). Notably, thyroid hormone–evoked changes in neu-
ronal function often involve glia-mediated actions (526). For
example, thyroid hormone up-regulates voltage-activated
sodium current in cultured postnatal hippocampal neurons
through T
3
-dependent secretion of basic fibroblast growth
factor from hippocampal astrocytes (528). Oligodendrocytes
are also responsive to T
3
and myelination is probably the best
characterized T
3
-mediated effect on glial cell function (529),
with T
3
stimulating myelin basic protein gene expression in
cultures of oligodendrocytes (530).
Cultured brain cells often exhibit substantial deiodinase ac-
tivity, D2 and/or D3, and thus the type and concentration of the
thyroid hormone applied in treatment groups should be con-
sidered and experiments planned accordingly. Frequent media
changes may be required to ensure stable iodothyronine con-
centration throughout the experiment. Adding tracer amounts
of radiolabeled T
4
or T
3
with subsequent analysis of the me-
tabolites in the media indicates their metabolic rate (173). In
coculture models, T
3
produced by D2 in H4 glioma cells affects
gene expression in neighboring SK-N-AS neuroblastoma cells
(173). In this system, concentrations of T
4
as low as 20 pM
evoked T
3
-mediated gene expression in the neuronal cells.
&
RECOMMENDATION 48b
Pituitary cell models respond to thyroid hormone in vitro
and can be used to study T
3
effects in the anterior pituitary
gland.
128 BIANCO ET AL.
Commentary. Pituitary cell lines are typically responsive
to thyroid hormone given the high TR expression in the gland
(101). The GH1 somatotroph cell line was among the first cell
models shown to respond to physiological levels of thyroid
hormone as monitored by cell growth and glucose utilization
(531). In the GH3 somatotroph cell line, exposure to T
3
in-
creases growth hormone expression in a TRb2-mediated
manner (532), while in GH-secreting GC cells, T
3
decreases D2
mRNA levels (143). The thyrotroph-derived cell line TaT1 also
responds to physiological levels of T
3
by decreasing TSHb and
D2 expression (259,533).
[J.2] Heart and cardiovascular system
Background. Development of the heart and transition of a
fetal to adult cardiac gene expression program is to a large
extent dependent on thyroid hormone (534). The heart re-
mains a major target organ of thyroid hormone in adult life
and cardiovascular effects are among the most pronounced
clinical manifestations of hypothyroidism and thyrotoxicosis
in man (535). Rodent hearts are equally responsive, and vir-
tually every aspect of cardiac biology is affected by thyroid
hormone, including electrophysiology, ion homeostasis, con-
traction, energy metabolism, and adrenergic signaling (402).
The sum of these effects is evident in the increases in heart rate
and rates of contraction and relaxation in the transition from
low to high thyroid-hormone states. Together with the reduc-
tion of peripheral resistance induced by thyroid hormone, this
results in the higher cardiac output required by the hyper-
metabolic organism. Ventricular growth is also stimulated by
thyroid hormone, almost doubling heart weight in the transi-
tion from hypothyroidism to thyrotoxicosis. However, thyroid
hormone has little direct effect on cardiomyocyte growth and
the ventricular hypertrophy is almost entirely in response to the
increase in cardiac workload induced by thyroid hormone
(537). Therefore, although the functional consequences of al-
tered thyroid-hormone levels are relatively straightforward
and well documented, separating the direct effects from the
secondary ones may be challenging. This even applies to those
cardiac genes that have been identified as direct targets of T
3
,
because they are in many cases transcriptionally coregulated by
factors that are themselves influenced by the thyroid status
(e.g., through load-dependent signal-transduction routes).
Understanding the mechanisms of thyroid-hormone action
in the heart and assessing cardiac T
3
activity is relevant,
particularly given the suspected role of impaired cardiac
thyroid-hormone signaling in the progression of heart disease
and the potential therapeutic use of thyroid hormone. This is
related to the low serum thyroid-hormone levels seen in heart
failure (i.e, the NTI syndrome) as well as to changes in cardiac
thyroid-hormone metabolism and TR expres sion in vario us
forms of heart disea se. The recommendations in this section
describe approaches and methods that are used to determine
the effect of the thyroid status on cardiac parameters. These
methodologies are also used to assess the role of cardiac
thyroid-hormone signaling in genetically modified mice,
such as TR knock-out models (451), and in models of path-
ological cardiac remodeling and heart failure. Models for
LV remodeling include the spontaneous ly hypert ensive rat
model (539); surgical induction of myocardial infarction by
transient or permanent ligation of the left descending coro-
nary artery (238,540); LV pr essure overload by surgical
banding of the aorta (transverse aortic constriction) (541–
544); and isoproterenol-induced LV hypertrophy in the
context of the D3 knock-out mouse model (64). Involvement
of thyroid-hormone signaling in right-ventricular (RV)
remodeling h as been studied in rat using the model of pul-
monary arterial hypertension induced by a single in jection of
monocrotaline (62, 63). The methods for inducing LV re-
modeling are used in rats and mice, but monocrotaline
cannot be used for inducing RV remodeling in mice.
&
RECOMMENDATION 49
The effects of thyroid hormone on cardiac function and mor-
phology can be assessed noninvasively by ultrasound or MRI.
Commentary. Noninvasive methods are used to study the
effects of thyroid hormone on cardiac function and morphol-
ogy (540,545,546). These methods are ideally suited for longi-
tudinal analyses of individual animals. Echocardiography is
the most widely used, allowing assessment of a range of
functional heart parameters (heart rate, stroke volume, cardiac
output, fractional shortening, ejection fraction) as well as ven-
tricular volumes and wall thickness during the cardiac cycle
(540,546). Ultrasonic devices with high spatial and temporal
resolution are required to image the mouse heart accurately
because of its small size and high heart rates (i.e., > 400 bpm).
With proper handling it is possible to analyze awake animals,
but echocardiography is typically done on lightly sedated an-
imals using isoflurane or tribromoethanol as anesthetics. These
compounds have minimal cardiodepressant effects, unlike
other commonly used anesthetics. Maintenance of body tem-
perature during the procedure is essential.
Certain aspects of fetal cardiac function can be assessed
using color Doppler–guided spectral Doppler ultrasound
(534). This specialized technique has been used in TR-mutant
mice to study the effects of disruption of T
3
action on func-
tional cardiac development in utero (534).
When the resolution of echocardiography is insufficient to
detect effects, MRI provides superior delineation of structures
throughout the cardiac cycle with high time resolution
(546,547). The same considerations with respect to anesthesia
and temperature control apply as for ultrasound analysis.
Image collection is triggered by the electrocardiogram (ECG)
and processing of acquired images allows reconstruction of
the heart through a complete cycle. From these data, func-
tional and structural cardiac parameters can be calculated
with greater accuracy.
&
RECOMMENDATION 50a
Analysis of the effects of thyroid hormone on heart rate
and blood pressure requires longitudinal analysis in free-
moving animals.
Commentary. Telemetry is required to monitor changes in
vital parameters accurately without interference from stress.
This method has been used in the analysis of thyroid-hormone
treatment of rats following myocardial infarction (540), in wild-
type mice (548), and in transgenic mice carrying TR mutations
(451,549). Heart rate, ECG, body temperature, arterial blood
pressure (550), and physical activity are simultaneously re-
corded by probes connected to a transmitter implanted into
the peritoneal cavity with a telemetry receiver located beneath
the cage. Continuous telemetry read out is possible for up to
INVESTIGATING THYROID HORMONE ECONOMY AND ACTION 129
3 months. Key issues are cost and skills required to perform
surgical implantation. A telemeter pre-implantation service is
available from suppliers including The Jackson Laboratories
(http://jaxmice.jax.org/preconditioned/surgical/telemetry.html)
and Charles River (www.criver.com/products-services/basic-
research/rodent-surgery/device-implants).
&
RECOMMENDATION 50b
The effects of thyroid hormone on LV contractile function
can be assessed by in vivo recording of LV pressures.
Commentary. Hemodynamic parameters are determined
in terminal e xperiments in anesthetized and ventilated mice
(238,551) or rats (540) by advancing a micro-tipped pressure
transducer or a combined pressure-volume catheter through
the right common carotid artery into the ascending aorta
and subsequently into the LV. The positive and negati ve
rates of pressure development and the sys tolic and end -
diastolic pressures may be used as sensitive indicators of
thyroid hormone–dependent changes in expression of con-
tractile proteins and proteins involved in intracellular cal-
cium homeostasis.
&
RECOMMENDATION 51a
Analysis of the effects of thyroid hormone on cardiac con-
tractile function requires ex vivo experiments using isolated
intact hearts or tissue preparations.
Commentary. The perfused heart preparation (Langen-
dorff method) is used to analyze a wide range of cardiac
function parameters, with the advantage of being able to
control experimental conditions (e.g., stimulus frequency, cor-
onary perfusion, ventricular volume, and pressure). A fluid-
filled PVC balloon is inserted into the LV cavity and attached to
a pressure sensor and data acquisition system, allowing re-
cording and manipulation of LV pressures and volumes. The
method can be applied to assess the cardiac effects of thyroid
hormone in both rat (552–554) and mouse models (555,556). An
additional advantage of this heart model is that it allows
analysis of the important effects of thyroid hormone on cardiac
substrate metabolism and mitochondrial function (552,556,
557). Labeled fatty acids and glucose are used to analyze oxi-
dative and glycolytic metabolism by standard techniques (
3
H-
labeled substrates) (552) or by NMR spectroscopy (
13
C-labeled
substrates) (556). This may be complemented by subsequent
tissue analysis of key metabolic enzymes and respiratory
function of isolated mitochondria (552,556).
More detailed analysis of the effects of thyroid hormone on
contractile and mechanical properties of cardiac tissue is
achieved using isolated trabeculae (558,559) or papillary mus-
cles (560). Force and shortening velocities of the electrically
stimulated muscles are recorded using a strain gauge. Papillary
preparations have also been used to record the effect of thyroid
hormone on electrophysiological properties (e.g., the shorten-
ing of action potential duration) (561). Thin preparations and/
or reduced temperatures are required to prevent tissue hypoxia
of the superfused preparations during repetitive stimulation
such as when determining force-frequency relationships. For
this reason small trabeculae are preferred, which are more
readily obtained from the RV than from the LV.
Right atrial preparations containing the sino-atrial node
can be used for recording isometric force and frequency of
spontaneous beating. This has been done to assess the effect of
aTRa1 mutation on the autonomic control of heart rate in
mice (549).
&
RECOMMENDATION 51b
The effects of thyroid hormone on cardiac Ca
2+
homeo-
stasis can be assessed in trabeculae or papillary muscles as
well as in isolated single cardiomyocytes.
Commentary. The positive inotropic and lusitropic effects
of thyroid hormone detected using trabeculae or papillary
muscles reflect changes in expression of myosin heavy chain
isoforms, as well as changes in expression of calcium-han-
dling proteins. The latter include the ryanodine Ca
2+
channel,
the sarcoplasmic/endoplasmic reticulum Ca
2+
-ATPase
(SERCA2a or Atp2a2) and its regulatory protein phospho-
lamban (536). The effect on beat-to-beat intracellular Ca
2+
fluxes can be assessed in trabeculae or papillary muscles using
Ca
2+
indicators (560,562,563). FURA2 is the most widely used
fluorescent indicator that allows determination of intracellu-
lar free Ca
2+
concentrations by a dual wavelength approach.
However, recording of Ca
2+
transients with millisecond time
resolution requires sophisticated high-speed optics. Dedi-
cated setups are available for combined analysis of contractile
properties and Ca
2+
fluxes (563).
The same experimental approach can be used to analyze
the effects of thyroid hormone on Ca
2+
homeostasis in indi-
vidual, electrically stimulated cardiomyocytes isolated from
mouse (564) or rat hearts (565,566). Analysis of single cultured
cells has obvious advantages, but the yield of viable, rod-
shaped cardiomyocytes is variable. It is particularly low in the
case of hypothyroid hearts, creating the risk of selection bias.
&
RECOMMENDATION 52a
The cardiac tissue response to thyroid hormone can be as-
sessed using histological methods and by studies of thyroid
hormone–responsive gene expression.
Commentary. The weight of the whole heart or individ-
ual ventricles should preferably be determined as a ratio to
tibia length, since body weight may vary considerably as a
function of the thyroid status. H&E staining of paraffin sec-
tions is typically used for measurement of cardiomyocyte
cross-sectional area (238,567). It should be noted that the
conditions of fixing cardiac tissue can affect cardiomyocyte
morphology and additional methods have been developed to
determine accurately myocyte cross-sectional area as well as
length (539). Masson’s trichrome staining of sections is used to
assess fibrosis in combination with cardiomyocyte morphol-
ogy (542), and the collagen volume fraction may be quantified
by picrosirius-red staining and polarized-light microscopy
(568). Expression of T
3
target genes, including Hcn2, Kcne1,
Kcnb1, Kcna1, Kcnq1, Thra, Thrb (534), Atp2a2, Myh6, Myh7
(62,238), and Nppa (238), is performed by RT-qPCR on mRNA
extracted either from whole heart or from individual cham-
bers or regions (see Section I.2 for technical considerations).
Expression of major cardiac myosin heavy chain isoforms
MHCa and MHCb (i.e., Myh6 and Myh7, respectively) can be
determined in tissue homogenates by Western blotting using
commercially available polyclonal (238) or monoclonal (542)
antibodies. IHC of T
3
-responsive genes is rarely done on the
assumption that all cardiomyocytes respond similarly to
130 BIANCO ET AL.
thyroid hormone or stress. However, recent studies have
shown an unexpected heterogeneity of regulated expression
of Myh7 (569,570), and this needs to be taken into account
when assessing effects of T
3
.
It is also important to realize that no cardiac protein is ex-
clusively regulated by thyroid hormone. The effectiveness of
T
3
treatment may be apparent from changes in expression of
the above mentioned genes—in particular the reciprocal
changes in Myh6 and Myh7 expression—but changes in ex-
pression of these genes in pathological situations does not
necessarily imply altered T
3
signaling. For instance, many of
the signal transduction routes involved in pathological ven-
tricular remodeling and heart failure also converge on the T
3
-
responsive genes, with effects opposite to those of T
3
.Asa
result, the phenotype of the failing heart resembles that of a
hypothyroid heart, at least for a number of proteins (535).
However, assessment of tissue T
3
levels (63,238) (see Section
B.2, Recommendation 8) and levels of TRs (571), or ideally,
determination of T
3
-dependent transcription (63,238) is re-
quired to establish changes in cardiac T
3
signaling.
&
RECOMMENDATION 52b
Contracting myocardium takes up plasmid DNA encoding
T
3
-responsive reporter genes, allowing cardiomyocyte-
specific T
3
-dependent transcription activity to be deter-
mined in vivo.
Commentary. Cardiomyocytes in the beating heart have
the unique capacity to take up plasmid DNA injected into the
ventricular wall. The mechanism of uptake is unknown, but
because no other cells present in the myocardium are trans-
fected, cardiomyocyte-specific regulation of fluorescent re-
porter genes can be studied in vivo. The same considerations
outlined for in vitro analysis of reporter genes apply with re-
spect to normalization and controls for off-target effects of
thyroid hormone or other interventions (see Section I).
However, analysis of T
3
-regulated transcription does not re-
quire co-transfection of a TR-expressing plasmid. During
thoracotomy surgery the free wall of the LV (mouse) or LV
and/or RV (rat) is injected with a small volume of plasmid
solution (Fig. 15). Typically, several thousand cardiomyocytes
are transfected, and expression lasts for several weeks. Ana-
lysis of reporter and normalization luciferase expression
(usually firefly and Renilla luciferase) is done on tissue ho-
mogenates. This methodology has been used to investigate
the T
3
-dependent regulation of the Myh7 promoter (572) and
to assess the T
3
-dependent transcriptional activity in rat and
mouse models of heart failure (63,238). Because in the latter
approach a minimal promoter containing only a T
3
RE as cis-
acting element is used, it yields the net result of all factors
potentially contributing to the altered T
3
signaling in the
cardiomyocyte (e.g., thyroid-hormone uptake, metabolism,
and expression of TRs and cofactors).
&
RECOMMENDATION 53
Primary cultures of rat and mouse cardiomyocytes, as well
as the cell line H9C2, can be used to study the effects of
thyroid hormone.
Commentary. Pri mary cultures of adult rat or cardio-
myocytes are o btai ned by Langendorff perfusion of the iso-
lated heart with a collagen-digesting solution. Rod-shaped,
viable cardiomyocytes are separated from permeabilized,
rounded cardiomyocytes and other cells and plated. The
FIG. 15. In vivo transfection of rat
cardiomyocytes by direct DNA in-
jection. The animal is anesthetized
with a mixture of N
2
O (0.2 L/min),
O
2
(0.2 L/min), and sevoflurane (2%–
3%), and the heart is exposed
through a right-lateral thoracotomy.
The free wall of the right and/or the
left ventricle is injected three to four
times each, delivering a total of 20 lg
of reporter plasmid(s)/ventricle in
100 lL of saline. In this example in-
jection of the right ventricle is shown,
with the 29-gauge needle bent at the
tip at an almost right angle to allow
easy injection of the thin right ventri-
cle wall. The thorax is then closed and
the animal is sacrificed 5 days later.
Expression of luciferase reporter and
normalization genes is determined in
ventricle homogenates. Courtesy of
Dr. Warner Simonides.
INVESTIGATING THYROID HORMONE ECONOMY AND ACTION 131
yield of viable cardiomyocytes varies, but it may be as high
as 80%. This depends primarily on the quality of the collage-
nase preparation used, and several batches of the enzyme may
need to be tested. Although primary cultures may last for more
than 1 week, the cardiomyocytes start to gradually dediffer-
entiate from the moment of isolation. Consequently, analyses
of the effects of altered thyroid status on cell parameters, such
as recording of action potentials (573) and Ca
2+
transients (564–
566), as well as in vitro effects of thyroid hormone (574), are
typically done on freshly prepared cultures.
Because of the limitations of adult primary cultures, most
studies of the effects of thyroid hormone on cardiomyocytes
in vitro are performed using primary neonatal rat (541,575–
577) or mouse (578) cultures. One- to three-day-old pups are
used and cells are plated following collagenase digestion of
minced ventricles. Neonatal primary cultures show sponta-
neous and synchronous contractions and cultures can be
maintained for more than 1 week. The interaction of thyroid
hormone and contractile activity can be analyzed either by
inhibiting spontaneous contractions (576) or by electrically
stimulating the cells (575). The cell line H9C2, derived from rat
embryonic heart tissue, is used as an alternative to primary
cardiomyocyte cultures and various aspects of thyroid-
hormone action and metabolism have been studied in H9C2
cells (579–581). Although this cell line has retained a number
of cardiomyocyte characteristics, it exhibits many of the key
properties of skeletal muscle. Notably, H9C2 myoblasts will
readily fuse to form multinucleated myotubes and respond to
acetylcholine stimulation. Care should therefore be taken
when extrapolating data to normal cardiomyocytes.
&
RECOMMENDATION 54
The effects of thyroid hormone on vascular function can be
assessed by analysis of arterial rings and cultured vascular
smooth muscle cells.
Commentary. The mechanism of thyroid hormone–
dependent reduction in systemic vascular resistance is
investigated by ex vivo determination of the vasomotor prop-
erties of rat arteries. Vascular rings are prepared from ex-
planted aortas or resistance arteries of hypothyroid, euthyroid,
or thyrotoxic rats and mounted in a myograph (582–585). The
contractile response to vasodilating or vasoconstricting agents
is recorded, allowing differentiation between changes in en-
dothelium-mediated effects and changes intrinsic to the vas-
cular smooth muscle cells (VSMCs). Primary cultures of
VSMCs are used to study the rapid effects of thyroid hormone
on VSMC relaxation (586) and the signal-transduction pro-
cesses involved (587).
[J.3] Intermediary metabolism and energy homeostasis
Background. All major metabolically relevant organs and
tissues (e.g., brain, adipose tissue, liver, and skeletal muscle)
are targeted by thyroid hormone. In most tissues, thyroid
hormone activates multiple metabolic pathways, leading to a
faster ATP turnover (ATP breakdown and synthesis) and
accelerated oxygen consumption (588,589). Energy transfer is
an inherently thermodynamically inefficient process; that is,
heat is an obligatory byproduct when energy is transferred (i)
from the oxidation of substrates into ATP and (ii) from ATP
into biological work. Thus, acceleration in ATP turnover leads
to heat production (thermogenesis), which is the accepted
pathway by which thyroid hormone activates thermogenesis.
Specifically, thyroid hormone accelerates turnover of a num-
ber of ATP-requiring ionic or substrate cycles including Na/K
ATPase, Ca ATPase, lipogenesis and lipolysis, and Cory cycle
(590). Both basal metabolic rate (i.e., energy spent to sustain
life in a resting state) and adaptive energy expenditure (e.g.,
cold-induced) are up-regulated by thyroid hormone. Direct
activation of the UCP1 gene transcription in BAT is the un-
derlying mechanism by which thyroid hormone accelerates
cold-induced thermogenesis (267,269,592); it is less clear that
thyroid hormone plays a role in diet-induced adaptive energy
expenditure (211,593). In addition to these direct actions in
energy homeostasis, thyroid hormone also acts in the CNS
(e.g., hypothalamus) influencing major homeostatic path-
ways; for example, sympathetic activity, appetite, and food
intake (594,595). Thus, studies of the metabolic effects of
thyroid hormone may also require strict control of food and
fluid intake, motor activity, respiration, and heart rate.
&
RECOMMENDATION 55a
Systemic thyrotoxicosis and hypothyroidism can be mod-
eled in rats; experimental endpoints to consider include
oxygen consumption (VO
2
), respiratory quotient (RQ),
feeding behavior, food and water intake, and movement/
activity.
Commentary. There is an approximately fourfold accel-
eration in the rate of energy expenditure during the transition
from hypothyroidism to thyrotoxicosis in rats (596,597). VO
2
is reduced to *50% once systemic hypothyroidism is
achieved (597). Administration of thyroid hormone rapidly
accelerates VO
2
in euthyroid or hypothyroid rats, with sig-
nificant changes observed as soon as 18–24 hours of the ad-
ministration (597,598). A significant drop in RQ of *10% is
observed within 2 days of the administration of thyroid hor-
mone to rats (599). The administration of T
3
to euthyroid rats
also results in rapid acceleration in VO
2
(*60%), which starts
at 24 hours and plateaus at 7 days (600).
The energetic cost of living for the mouse is about twofold
higher than for the rat (601). Thus, substantial differences exist
between mice and rats with respect to their metabolic re-
sponsiveness to thyroid hormone. Hypothyroid mice do not
exhibit a decrease in VO
2
at room temperature, only after
acclimatization at thermoneutrality (211). However, the total
daily energy expenditure is decreased in the hypothyroid
mouse acclimatized at room temperature, a parameter that
takes into account the VO
2
and the RQ. The sustained total
daily energy expenditure in the hypothyroid mouse is due to
enhanced sympathetic stimulation of BAT (211,602). This in-
dicates that the reduction in thyroid hormone signaling is
compensated for by an increase in sympathetic activity, sus-
taining VO
2
at room temperature (211).
Indirect calorimetry can be studied over 30–120 minutes
with reliable results (597,598,603,604). Stress might be a sig-
nificant factor in these short-term studies and care should be
taken to avoid or at least control for such a variable. In some
settings animals are anesthetized to minimize movements
and stress (597), but anesthetic agents can interfere with
thermoregulation and energy homeostasis. A state-of-the-art
methodology involves admitting mice to an integrated con-
tinuous monitoring system that can house animals for days or
132 BIANCO ET AL.
weeks at controlled temperature (Fig. 16) (232). This mini-
mizes the effects of stress in addition to allowing for acquisi-
tion of 24 hour data profiles. Another major advantage is
obtaining data points during the day (which reflect basal
metabolic rate) and night (which include acceleration of the
metabolic rate caused by feeding).
Ideally, VO
2
is expressed as a function of lean body mass
de termined at or around the time of indi rect calori metry. Fat
and lean body masses can be measured by destructive car-
cass composition analysis or by more recent noninvasive
imaging techniques including dual energy x-ray absorptio-
metry or quantitative MRI. These methods have been com-
pared and may yield differing results, further indicating the
importance of careful e xperimental design and the use of
complementary methods for analysis of metabolic pheno-
types (606,607).
FIG. 16. Use of the Comprehensive Laboratory Animal Monitoring System to perform continuous indirect calorimetry in
mice. (A) Two independent metabolic cages that are connected to a computer for recording and data analyses; (B) VO
2
during
a 24 hour time period showing the nocturnal (shaded area) increase in metabolism; (C) respiratory quotient (RQ) during the
same period of time, depicting the nocturnal (shaded area) increase in RQ when animals are kept on a carbohydrate-enriched
diet; (D) dramatic increase in VO
2
during 48 hours cold exposure (shaded area); (E) decreased VO
2
in hypothyroid mice that
were kept on 0.1% MMI for 60 days; (F) RQ in the animals shown in (E). Courtesy of Drs. Antonio Bianco and Tatiana
Fonseca.
INVESTIGATING THYROID HORMONE ECONOMY AND ACTION 133
Core temperature in rodents can be measured via a rectal
thermal probe connected to a digital thermometer (608).
However, state-of-the-art technology involves surgically im-
planting radio transmitters into the abdominal cavity to
monitor core temperature (609,610). Administration of anti-
thyroid drugs to rats results in rapid reduction of core tem-
perature that is evident after 3 days of treatment; by the end of
the first week of treatment, there is an approximately 0.5C
reduction in core temperature that is sustained during all day
and night (609). At the same time, rats treated with thyroid
hormone for 1 week exhibit an increase of about 1C in core
temperature as assessed with a rectal probe (611).
&
RECOMMENDATION 55b
The effects of systemic hypothyroidism or thyrotoxicosis
on a particular tissue can be studied via tissue or cell iso-
lation from appropriately treated animals.
Commentary. Thyroid hormone is well known to influ-
ence the metabolic rate of intact organisms; however, the
distribution of the effect among various tissues is not as well
understood. The liver is clearly established as an important
mediator of the metabolic effects of thyroid hormone based on
a series of studies utilizing hepatocytes isolated from rats with
experimentally induced hypothyroidism or thyrotoxicosis
(612), reviewed by Harper and Brand (613). Hepatocyte oxy-
gen consumption was positively correlated with thyroid
hormone status of the animal; analysis of respiration indi-
cates that basal mitochondrial proton leak is also positively
correlated with oxygen consumption (614). At the same time,
other cellular mechanisms differ in their response to hypo-
thyroidism and thyrotoxicosis: in hypothyroid cells, non-
mitochondrial oxygen consumption decreases, whereas in
thyrotoxic cells reactions involving ATP turnover are in-
creased (613). It should be noted that because the hepatocytes
were made hypothyroid or thyrotoxic in vivo (while still part
of an intact liver in a pharmacologically treated animal), these
effects on respiration cannot be assumed to be directly related
to effects of T
3
on hepatocytes but may include hepatic effects
caused by second messengers arising from extrahepatic
sources such as the sympathetic nervous system (SNS).
Historically, energy expenditure of isolated cells has been
quantified by measuring oxygen consumption using a Clark
electrode or similar apparatus (234). Alternatively, oxygen
consumption and media acidification can be measured non-
destructively in 24- or 96-well format using modern instru-
ments based on solid state fluorophore/optical sensors
(Fig. 17) (615,616).
&
RECOMMENDATION 55c
To isolate the direct effects of thyroid hormone on cellular
energy expenditure and metabolic pathways on a given cell
type, in vitro studies should be performed with thyroid
hormone treatment of cultured cells.
Commentary. Studies in which cultured cells are exposed
to varying concentrations of thyroid hormones in the medium
exclude the possibility of second messengers arising from
FIG. 17. Measuring O
2
consumption and
the rate of medium acidification in cultured
cells. Extracellular flux analysis is per-
formed using XF Analyzers (Seahorse
Bioscience, Billerica, MA). Cells are plated in
monolayer 24- or 96-well format. Oxygen
consumption rate (OCR) and extracellular
acidification rate (ECAR) are measured via
chemiluminescent sensors applied to dis-
posable cartridge-based probes that are
lowered over the cells and a few microliters
of media (lower picture), creating a tran-
siently sealed micro-environment. As oxy-
gen is depleted from the media and protons
accumulate in the media, the device plots
the concentration of O
2
and pH in real time
(upper picture). The OCR and ECAR are
calculated as the slopes of the O
2
and pH
versus time curves. Modified with permis-
sion from a web posting by Seahorse
Bioscience; courtesy of Dr. Brian Kim.
134 BIANCO ET AL.
other tissues, and thus may reveal ‘direct’ effects of thyroid
hormone on a given cell type. For example, gene expression
analysis of hepatocellular carcinoma cells (HepG2) exposed to
pharmacologic concentrations of T
3
has demonstrated chan-
ges in metabolically relevant genes (617). Alternatively, pri-
mary hepatocyte cultures from rats or mice may be used for
such studies because they represent excellent models of thy-
roid hormone effects on the liver (618,619). In addition, D3
activity has been shown to be inversely correlated with oxy-
gen consumption in cultured cells, suggesting a direct effect of
T
3
on energy expenditure (though T
3
concentrations were not
measured in the intracellular compartment) (63).
&
RECOMMENDATION 55d
Studies utilizing isolated mitochondria can be used to in-
vestigate the effects of T
3
on mitochondrial biology.
Commentary. The effects of systemic, pharmacologic al-
teration of thyroid status in rats on hepatic mitochondria have
been investigated following their isolation from treated ani-
mals; mitochondrial proton leak and the respiratory chain
were identified as important sites of T
3
control (620,621).
While such studies present advantages in that mitochondria
can been interrogated apart from other cellular mechanisms,
the respiratory phenotype is different from that observed in
isolated cells because the physiologic context is lost; for ex-
ample, the increase in ATP-consuming reactions induced by
thyrotoxicosis is not seen given the lack of extramitochondrial
ATP-consuming pathways (613).
&
RECOMMENDATION 56a
Acute and chronic responsiveness to cold exposure (4C–
5C) can be used to study the thermogenic effects of thyroid
hormone.
Commentary. Defending core temperature during cold
exposure depends on a series of adaptive mechanisms mostly
initiated by the SNS. However, because of modifications in
the adrenergic signal transduction system, hypothyroid ani-
mals respond much less to catecholamines, and the opposite is
observed in thyrotoxic animals (622–624). Thus, hypothyroid
rodents exhibit profound hypothermia and succumb in a
matter of hours when exposed to cold (4C–5C) (625,626). By
2 hours of being moved to cold all rodents exhibit a transient
drop in rectal temperature (about 1C) but recover to baseline
values by 4 hours of continued cold exposure (610). In hy-
pothyroid rats and mice, rectal temperature continues to de-
crease, and to avoid the animals’ death the experiment should
be interrupted at core temperatures of about 30C.
&
RECOMMENDATION 56b
BAT thermogenesis is positively regulated by thyroid
hormone and can be studied in the interscapular BAT
(iBAT) pad during infusion with adrenergic agonists in rats
or mice.
Commentary. BAT is an important site of thermogenesis
and has been used extensively as a model to study thyroid
hormone signaling and its synergism with the SNS. Heat
production in the BAT is initiated by SNS stimulation. Thus,
the ultimate assessment of BAT thermogenesis should include
measurement of BAT temperature in response to adrenergic
stimulation. This can be obtained by measuring the thermal
response of iBAT to stimulation with an adrenergic agonist
(e.g., norepinephrine) in a dose- and time-dependent fashion
(627). The setup includes surgical placement of a small
thermistor under the iBAT (Fig. 18). Due to the small body
size, mice should be kept on a thermal bed set to 30C at all
times (311). The jugular or femoral veins are cannulated and
connected to a pump and used as ports for drug infusion. A
second thermistor should be inserted in the rectum to monitor
core temperature. iBAT thermal response can be observed
within minutes of the start of the infusion. The increase in
iBAT temperature is usually 1C–2C in the first hour and can
reach up to 4C depending on whether the animals were
previously fed a high-fat diet or chronically exposed to cold
(593). The core temperature should increase less markedly
and at later times as an indication that the BAT is warming up
the body and not the other way around. This setup allows for
direct assessment of the thyroid hormone effects in the BAT,
bypassing indirect effects. For example, the hypothyroid iBAT
is only minimally responsive to an infusion with norepi-
nephrine (627,629).
&
RECOMMENDATION 57a
Studies of metabolic effects of thyroid hormone may need
to be performed under conditions of thermoneutrality; that
is, the ambient temperature at which core temperature is
maintained by obligatory thermogenesis (thermogenesis
produced by the animals at its basal metabolic rate).
Commentary. Room temperature (21C) is a significant
thermal stress for small rodents, especially mice. Thus, even in
animals acclimatized to room temperature, the metabolic ef-
fects of thyroid hormone are influenced by a significant syn-
ergism with sympathetic activity. Acclimatization at 30C
minimizes the sympathetic activity and thus interference in
metabolism. Most sympathetic effects on metabolism are
minimized within the first 24–48 hours of acclimatization at
30C (232). However, it may take up to several weeks for a full
sympathetic shut down and investigators may need to de-
termine this experimentally depending on the parameters to
be studied (232,608). This is particularly important if the
studies involve hypothyroid or thyrotoxic animals given that
the activity of the SNS is inversely correlated with thyroid
status (630,631); it is increased during hypothyroidism and
minimized during thyrotoxicosis (632,633).
&
RECOMMENDATION 57b
The use of diet-induced obesity as a model to study the
thermogenic effect of thyroid hormone must take into
account additional thyroid hormone-dependent variables;
for example, growth hormone secretion and linear
growth, appetite, sympathetic activity, and environment
temperature.
Commentary. BAT is also a main site of diet-induced
thermogenesis in small rodents, which can be triggered by
feeding a high-fat diet (634). However, contrary to cold-
induced thermogenesis, the role played by thyroid hormone
in diet-induced thermogenesis is less clear. Hypothyroid rats
or mice are not obese when kept on a chow diet (211,593),
but when they are placed on a high-fat diet, the response is
different between rats and mice and depends on the
INVESTIGATING THYROID HORMONE ECONOMY AND ACTION 135
environmental temperature (211,593). At room temperature,
hypothyroid rats placed on a high-fat diet become just as
obese as euthyroid control rats (593), whereas mice are pro-
tected against diet-induced obesity (211). Only at thermo-
neutrality do hypothyroid mice become obese when placed
on a high-fat diet (211). For these types of studies, commer-
cially available diets are preferred for consistency and may
contain anywhere from 20% to 50% fat. Such diets contain
approximately 4.73 kcal/g in which 45% of energy is derived
from fat and can be compared to responses in mice fed a
standard diet consisting of 3.85 kcal/g in which 10% of energy
is derived from fat (608). In general, animals are exposed to
different dietary regimens for periods that vary between 30
and 60 days.
&
RECOMMENDATION 58
Thyroid hormone–dependent metabolic processes in adi-
pose tissue can be investigated in tissue preparations, in
freshly isolated white or brown adipocytes, and in white or
brown adipocytes differentiated in vitro.
Commentary. The structural or biochemical analysis of
fat pads in hypothyroid or thyrotoxic mice or rats constitutes
a basic approach to study the effects of thyroid hormone
in adipose tissue. Thyroid hormone affects the weight of
specific fat pad deposits (e.g., epididymal, inguinal, and in-
terscapular) (635), which should always be expressed as a
function of total body weight or femoral length (464). Lipo-
genesis is stimulated by thyroid hormone in white and brown
adipose tissues (237,268,637). At the same time, thyroid hor-
mone also promotes lipolysis via amplification of the adren-
ergic signaling pathway. The net balance of these antagonistic
pathways is a reduction in adiposity, reaching *50% in 6
days (600). Although one would intuitively speculate that in
hypothyroid animals there should be an increase in adiposity,
body fat in hypothyroid rats remains unaffected (593),
whereas it is decreased in hypothyroid mice (211).
In the white adipose tissue and BAT thyroid hormone
stimulates lipogenesis and lipolysis; only in BAT does T
3
stimulate thermogenesis. UCP-1 gene expression (mRNA and
protein) is the critical marker for BAT thermogenesis, and it
is highly responsive to thyroid hormone (267,269,592,638).
Other thyroid hormone–responsive pathways exist in the
adipose tissue that can be assessed by measuring the Vmax of
key rate limiting enzymes (268,269,640). Examples include
the assays of acetyl CoA carboxylase activity by an NADH-
coupled assay (237), malonyl CoA decarboxylase activity
using a carnitine acetyltransferase-linked assay (641), and the
activity of a-GPD in mitochondria preparations using a L-a-
glycerophosphate as substrate (464), although investigators
should be aware that a-GPD activity may not be thyroid
hormone dependent in all mammalian species. Lipogenesis is
particularly sensitive to thyroid hormone and can be assessed
by measuring de novo fatty acid synthesis in fat pad depots
after intraperitoneal or intravenous injection of
3
H
2
O fol-
lowed by lipid extraction at different time points (600,642).
Isolated white or brown adipocytes prepared from colla-
genase digestion of specific fat pads can be studied in
FIG. 18. Study of the interscapular
brown adipose tissue (iBAT) thermal
response to norepinephrine (NE) infu-
sion. (A, B) A rat (or mouse) is an-
esthetized, and the iBAT pad is
exposed through a surgical incision. A
thermistor is placed under the iBAT
pad and secured with a stitch; a rectal
thermistor is inserted in the colon for
measurement of core temperature (not
shown). The right jugular vein is can-
nulated and connected to an infusion
pump for infusion of catecholamines
or other molecules. Temperatures are
measured continuously before and
during infusion. Courtesy of Dr. Mir-
iam O. Ribeiro. (C) Plasma NE levels
during infusion in intact and Tx rats;
(D) iBAT temperature during infusion
in intact and Tx rats. Infusion lasted
for 60 minutes, and the temperature
data points indicate the difference be-
tween baseline and maximum peak
achieved during infusion. Reproduced
with permission from Ribeiro et al.
(627).
136 BIANCO ET AL.
suspension for a few hours only, given that prolonged incu-
bations are hardly stable (237,622,643). An advantage of this
approach is that cellular metabolism can be studied under
defined conditions of substrate availability; for example, low/
high glucose, fatty acids, and carnitine (644). The results ob-
tained reflect the state of these cells immediately before being
harvested. Specific metabolic pathways can be studied by
measuring the Vmax of key rate-limiting enzymes under basal
conditions and in response to provocative agents such as
norepinephrine, forskolin (activating adenylate cyclase), di-
butyryl cAMP (a phosphodiesterase-resistant cAMP ana-
logue), terbutaline (b
2
-selective adrenergic agonist), and
dobutamine (b
1
-selective adrenergic agonist) (464). Examples
include measuring the rate of energy expenditure (VO
2
), li-
pogenesis, lipolysis, or beta-oxidation, all thyroid hormone-
sensitive pathways (237). Lipogenesis can also be studied
after a pulse with
3
H
2
O and lipid extraction (237). Lipolysis is
also studied by measuring catecholamine-stimulated glycerol
release (645–647).
Isolated adipocytes have limited in vitro responsiveness to
thyroid hormone (237,648,649). In contrast, primary cultures
of white or brown adipocytes are responsive to thyroid hor-
mone and thus constitute an advantageous alternative system
(69,651). Obviously primary cultures of adipocytes do not
reflect the metabolic status of an animal, but they have the
advantage of generating cells that are stable and thus suitable
for prolonged experiments. Mature adipocytes are obtained
after white or brown pre-adipocytes are pushed in vitro to-
wards differentiation by exposure to different cocktails and
strategies during an 8–12 day period (652). Lines of white or
brown pre-adipocytes have also been established and used
successfully (653).
&
RECOMMENDATION 59a
Simultaneous induction of lipogenesis and fatty acid oxi-
dation are important effects of thyroid hormone in the liver.
Studying expression of key rate-limiting enzymes as well as
estimating the rates of both of these processes in animal and
cell models can be used to gauge the extent of thyroid
hormone signaling in the liver.
Commentary. With the second highest density of TRs in
the body, the liver is highly responsive to thyroid hormone
(654). Eighteen thyroid hormone–responsive proteins were
identified several years ago in the analysis of a translational
assay of hepatic mRNA of euthyroid and hypothyroid
(655). As with the fat tissue, thyroid hormone stimulates
lipogenesis in the liver by activating the expression of key
enzymes involved in the synthesis of fatty acids (596), such
as malic enzyme (656), glucose-6P-dehydrogenase (both
enzymes generating NADPH), fatty acid synthase, and acetyl-
CoA carboxylase (657–659). At the same time, thyroid hor-
mone rapidly accelerates fatty acid oxidation in the liver (660).
There is an approximately fivefold reduction in fatty acid
oxidation in the isolated hepatocytes of hypothyroid rats. In
contrast, the administration of thyroid hormone to euthyroid
rats accelerates the rate of fatty acid oxidation (661). This ex-
plains in part the reduction in RQ seen during the transition
between hypothyroidism and thyrotoxicosis (599). However,
the net result of thyroid hormone action in liver triglycerides
tends to be neutral given the combined effects on lipolysis,
lipogenesis, and fatty acid oxidation (211).
&
RECOMMENDATION 59b
Serum cholesterol levels and liver cholesterol content can
be used as biological markers of the metabolic effects of
thyroid hormone in the liver.
Commentary. Hypothyroid rats typically have high
plasma cholesterol with normal, reduced, or marginally ele-
vated triglycerides (662–665). The hypercholesterolemia is
largely caused by an increase in cholesterol concentration in
low-density lipoprotein (LDL) that results from decreased
receptor-mediated catabolism of the lipoprotein, primarily in
the liver. This is intensified by feeding rodents with a high
cholesterol diet. A widely utilized model is feeding a high-fat
diet, such as 10% corn oil containing 2% cholesterol (choles-
terol diet) or 10% corn oil, 2% cholesterol, and 0.5% cholic acid
(cholic acid diet). Hypothyroid rats kept on such diets exhibit
marked hypercholesterolemia (665) and hepatic secretion of
cholesteryl ester and apoE-rich very low density lipoprotein
and LDL (666,667). In contrast, thyrotoxic rats have reduced
cholesterol levels in association with reduced LDL turnover.
Thyroid hormone administration affects cholesterol metabo-
lism as early as 1 week after treatment (668). This has fostered
considerable interest in the potential for thyroid hormone
analogs in the treatment of hyperlipidemic disorders
(310,315,669,670).
[J.4] Skeletal muscle
Background. The contractile and metabolic properties of
skeletal muscles are determined by the properties of the dif-
ferent fiber types that make up the muscle. Groups of fibers
innervated by a single motor neuron form a motor unit. The
type of innervation and thyroid hormone are the principal
modulators of the phenotype of muscle fibers, both during
development and in the adult (671). The dynamic interaction
between these modulators, together with the developmental
origin of the fibers, results in various phenotypes ranging
from slow-twitch, oxidative fibers (Type I) to fast-twitch,
glycolytic fibers (Type IIB) (671). Skeletal muscle fibers are
typically identified by the major type of MHC that is ex-
pressed; that is, MHC I, IIa, IIx/d, and IIb (note that MHC I is
identical to cardiac MHC b, but that the other MHC isoforms
are not expressed in heart). These isoforms have increasing
catalytic turnover rates, with MHC IIb being five times faster
than MHC I, and largely determine the rate of contraction.
Multiple isoforms also exist for other key proteins that make
up the muscle fiber, such as the myosin light chains (MLCs)
and the sarcoplasmic/endoplasmic reticulum Ca
2+
-ATPases
SERCA2a (Atp2A2) and SERCA1a (Atp2A1). These SERCA
isoforms are responsible for the re-uptake of cytosolic Ca
2+
,
determining the rate of relaxation, and are typically associated
with slow (SERCA2a) or fast fibers (SERCA1a) (672). Apart
from obvious differences in mitochondrial density between
primarily oxidative and glycolytic fibers, marked qualitative
differences also exist in composition and function of the mi-
tochondria in these fibers (673).
The fiber composition of different muscles in the body
varies considerably, matching their specific tasks. Develop-
ment of slow characteristics is dependent on innervation by
slow motor neurons, whereas development of fast character-
istics is much less dependent on neural input (674). Expres-
sion of the fast phenotype is, however, co-regulated by
INVESTIGATING THYROID HORMONE ECONOMY AND ACTION 137
thyroid hormone (671,674). For example, the postnatal ex-
pression of SERCA1a is completely dependent on thyroid
hormone (675,676) and the transition from fetal to adult fast
MHC isoforms is delayed in hypothyroidism (677–679). In the
adult animal this responsiveness of fast muscles is greatly
diminished, in contrast to slow muscles, which are highly
sensitive to changes in the thyroid status. The reason for this is
that, generally speaking, thyroid hormone opposes the effect
of slow-type innervation; that is, it stimulates expression of
fast characteristics, while suppressing slow characteristics,
and it increases the overall ATP-generating capacity of the
muscle. Fast muscle fibers are energetically less efficient and
the increase in speed induced by thyroid hormone is associ-
ated with higher energy turnover and concomitant heat pro-
duction. Because of the total mass of skeletal muscle, these
effects of thyroid hormone have an impact on whole body
substrate metabolism and thermogenesis (671).
The basis of muscle plasticity is the fairly coordinated
regulation of expression of genes governing contractile
properties and energy metabolism in response to changes in
external cues, such as use or thyroid status. Although the
overall effects of thyroid hormone on muscle physiology are
relatively straightforward, the effects on gene expression at
the level of individual muscle fibers are complex and may be
diametrically opposed, depending on the context (676,680).
For instance, expression of MHCIIa is strongly stimulated by
thyroid hormone in predominantly slow muscles, but equally
strongly repressed in fast muscles (674,680). Furthermore,
expression of both SERCA1a and SERCA2a is transcription-
ally stimulated by T
3
. However, in thyrotoxic slow muscle the
expression of SERCA2a is shut off in the majority of fibers and
replaced by SERCA1a (676), but at the same time they are co-
expressed at high levels in other fibers. This primarily reflects
subtle differences in the various motor units that make up the
muscle. Consequently, whole muscle analysis of gene expres-
sion may not provide an accurate picture of the effects of thy-
roid hormone, particularly when working with mixed fiber
type muscles. Such studies should therefore include analyses of
tissue sections to determine effects in individual fibers.
In larger mammals, all skeletal muscles contain a variable
mix of fast and slow fibers. However, in rodents some hind
limb muscles are primarily composed of slow type I and in-
termediate type IIA fibers, such as the slow-twitch soleus
(SOL) muscle, or fast type IIB fibers, such as the fast-twitch
extensor digitorum longus (EDL) muscle (681,682). These
muscle are therefore often used as models to study the effects
of thyroid hormone on slow and fast muscle properties.
&
RECOMMENDATION 60
The effects of thyroid hormone on muscle energy metabo-
lism and contractile properties can be assessed by analysis
of perfused hind limb preparations.
Commentary. In the pump-perfused muscle model a
single hind limb of the rat is perfused following catheteriza-
tion of the femoral artery and vein (683). Contraction of a
group of muscles or a single muscle is triggered either by
stimulating the innervating nerve or by using electrodes
placed on the muscle. Contraction parameters are recorded
using a force-transducer connected to the Achilles tendon.
Control of the rate and composition of the perfusion medium
allows manipulation of substrate and oxygen delivery. Ana-
lysis of the venous perfusate yields information on the met-
abolic response during different stimulus protocols and
allows calculation of the efficiency of contraction. Blood flow
distribution over the various perfused muscles can be deter-
mined using radiolabeled microspheres, and histochemical
and biochemical analyses can complement the analyses of
muscle function (684). Although there are some limitations,
notably the absence of normal vasomotor regulation (683),
this model best approximates the in vivo situation of active
skeletal muscles and has been used to study the functional
and metabolic consequences of thyrotoxicosis (685) and hy-
pothyroidism (684,686). This methodology has so far not been
described for mouse models.
&
RECOMMENDATION 61
The interaction between thyroid hormone and neuromus-
cular activity in regulating muscle properties can be stud-
ied in vivo by direct electrical stimulation of muscles.
Commentary. Slow-twitch muscles like the SOL muscle
are typically activated by motor neurons generating long
trains of low-frequency impulses, whereas fast-twitch mus-
cles are activated intermittently by short bursts of high
frequency impulses. The counteracting effects of thyroid hor-
mone and slow-type innervation on muscle gene expression
(674) can be studied by imposing a chronic low-frequency
stimulation pattern on fast-twitch hind limb muscles. In this
model the peroneal nerve of the hind limb is continuously
stimulated in vivo at 10 Hz for up to 20 days through implanted
electrodes. Gene expression and fiber type composition of fast-
twitch tibialis anterior and EDL muscles from the stimulated
and contralateral hind limb can then be compared. This
method has been used to assess the combined effects of thyroid
hormone and slow-type innervation on the expression of
myogenic regulatory factors (687) and MHC isoforms (688).
&
RECOMMENDATION 62
The effects of thyroid hormone on contractile properties
can be assessed by ex vivo analysis of intact muscles and
isolated muscle fibers.
Commentary. Detailed analysis of the contractile prop-
erties of rat and mouse skeletal muscles requires ex vivo ana-
lyses. The slow-twitch SOL muscle has typically been used in
studies involving thyroid hormone (689–692) because of the
greater responsiveness when compared to the fast-twitch EDL
(693). Although contractile properties can be recorded with
the muscle left in situ, analysis of isolated muscles in a stim-
ulation chamber allows for greater experimental flexibility
(690). The EDL and SOL muscles are relatively thin and su-
perfusion provides sufficient delivery of oxygen and sub-
strates, although reduced temperatures (21C–30C) may be
required to prevent hypoxia during prolonged stimulation.
Contralateral muscles are kept in oxygenated superfusion
medium to serve as controls in subsequent biochemical ana-
lyses. Typical parameters analyzed include twitch and tetanic
contractile response, force-frequency relationships, post-rest
potentiation of force development, and fatigue-recovery re-
sponses. Dedicated equipment and software is now available
for these kinds of analyses (694).
Although technically challenging, single muscle fibers can
be isolated from skeletal muscles such as the SOL and EDL
138 BIANCO ET AL.
muscles (692,695) and the diaphragm (696). In the latter,
thyroid hormone also induces a shift in fiber phenotype type
from slow to fast (697) and ex vivo analysis of the contractile
properties of this muscle requires single fibers or bundles of
fibers (698). Isolated fibers are chemically skinned and
mounted between a force transducer and a motor, allowing
measurement of isometric and isotonic contractions. Con-
traction or relaxation is achieved by immersion of the fibers in
solutions containing high or low concentrations of calcium,
respectively. Subsequent single-fiber analysis of contractile
protein composition allows for a detailed assessment of the
effects of thyroid hormone on fiber phenotype and contractile
properties (692).
&
RECOMMENDATION 63
The skeletal muscle tissue response to thyroid hormone can
be assessed by analysis of gene expression and enzyme
assays.
Commentary. Analysis of muscle gene expression is
performed by RT-qPCR (92,699,700) (see Section I.2,
Recommendation 35 for technical considerations) and by
analysis of protein levels using polyacrylamide gel electro-
phoresis, followed by quantification of stained bands or Wes-
tern blot analysis where appropriate. The MHC isoforms, as
well as the MLC isoforms, can be separated by high-resolution
gel electrophoresis. Silve r staining and subsequent densit o-
metric analysis allows accurate determination of the relative
expression levels of the various isoforms in samples as s mall
as single fibers (679,692,699,701,702). Antibodies for Western
analysis are available for a large number of skeletal muscle
proteins encoded by genes which are directly or indirectly
regulated by thyroid hormone, including MyoD, myogenin,
MHC i soforms, SERCA1a, SERCA2a, myoglobin, PGC-1a,
RYR, and Na
+
-K
+
-ATPase subunits ( 92,691,701,703,704).
The sarcoplasmic reticulum (SR) is responsible for up to
50% of the energy consumption in active muscle. The effect of
thyroid hormone on this ATPase activity can be assessed
spectrophotometrically in homogenates of whole muscles or
single fibers (701,705). SR membrane fragments in muscle
homogenates reseal into vesicles that retain the ability to take
up and store Ca
2 +
. Using assay conditions that selectively
stimulate this capacity, it is possible to determine the ATPase-
coupled Ca
2 +
-uptake rate and maximal Ca
2 +
-storage capac-
ity (675,704,705). The latter provides an estimate of the degree
of proliferation of the SR membrane network, which is in part
dependent on the thyroid status. The Ca
2 +
-ATPase and Ca
2 +
-
uptake assays determine the total SERCA activities, but do
not allow discrimination between the SERCA1a and SER-
CA2a isoforms.
Standard assays are used to determine activities of en-
zymes involved in aerobic and anaerobic energy metabolism
such as cytochrome c oxidase, citrate synthase, succinate de-
hydrogenase, lactate dehydrogenase, and creatine kinase
(675,703). Mitochondria can be isolated by differential cen-
trifugation for measurements of respiratory rate, membrane
potential and other enzymatic assays (706,707).
&
RECOMMENDATION 64
Assessment of the effects of thyroid hormone on the fiber
type composition of skeletal muscle requires histochemical
analysis of tissue sections.
Commentary. Skeletal muscle fiber types are primarily
distinguished on the basis of the MHC isoform expressed; i.e.,
the slow MHC I and fast MHC IIa, IIx/d, or IIb isoforms.
Antibodies discriminating between these isoforms have been
described (708) and although some caution is warranted
when using currently available commercial antibodies (709),
multicolor immunofluorescence analysis has been described
for the simultaneous assessment of expression of all
MHC isoforms in rat and mouse skeletal muscle using these
antibodies (710). More typically, a histochemical method
for myofibrillar ATPase expression is widely used which
distinguishes between Type I fibers (MHC I), Type IIB fibers
(MHC IIb), and intermediate fibers (MHC IIa, MHC IIx/d)
(684,692,700,702,711). Muscles are clamped at their approxi-
mate in situ length and frozen in liquid N
2
-cooled freon or
isopentane. Series of cryosections are then cut for various
histochemical analyses. It should be noted that not all fibers
necessarily run the length of the muscle. This is for example
the case in SOL muscle, which should preferably be sectioned
at the motor point to include all fibers (682,702).
Additional analyses of fiber-specific gene expression con-
cern the expression and activity of the SR Ca
2 +
-ATPase
enzymes. Serial cryosections can be used to correlate MHC-
based fiber typing with immunohistochemistry of SERCA1a
and SERCA2a (676,682), as well as with isoform-specific
mRNAs by in situ hybridization (676). Fiber-specific differ-
ences in total SR Ca
2 +
-ATPase activity can also be determined
histochemically (676). The effects of thyroid-hormone treat-
ment on fiber characteristics are dynamic and in some cases
transient over a period of several days (676). This should be
taken into account when studying animals treated for less
than 14 days.
&
RECOMMENDATION 65
Primary cultures of skeletal muscle cells as well as various
cell lines can be used to study thyroid hormone–regulated
gene expression and its interaction with contractile activity.
Commentary. Pri mary cultures of mu scle myoblasts can
be obtained from neonatal rat or mouse skeletal muscles
(92,712–715). Depending on the composition of the culture
medium, myoblasts can be maintained in a proliferative
state, or they can be induced to fuse and form spontane-
ously contracting myotubes. The interactio n of thyroid
hormones and contractile activity on myotube properties
can be studied by inhibiting contractions (715) or by direct
electrical stimul ation of the myotubes in culture (712). Al-
though primary cultures prov ide a versatile model to study
different aspects of muscle gene regulation, it should be
noted that myotubes do not fully develop adult phenotypic
characteristics.
The mouse C2C12 (704,716–718) and the rat L6 (704,717,
719–721) myoblast cell lines are used as alternatives for pri-
mary cultures in studies of thyroid-hormone action. Standard
techniques for cell transfection can be used to analyze regu-
lation of gene-expression (see Section I.4). Subclones of the L6
cell line with slightly different properties have been isolated
by various groups (704,721). Notably, the L6
AM
subclone (721)
shows increased responsiveness to thyroid hormone in com-
parison to the original L6 cell line. Similar to primary cultures,
confluent myoblast cultures of both cell lines can be induced
to form myotubes, but only C2C12 myotubes can be induced
INVESTIGATING THYROID HORMONE ECONOMY AND ACTION 139
to contract by electrical stimulation. Methods have been de-
veloped to stimulate C2C12 myotubes chronically or inter-
mittently for up to 4 days using stimulus electrodes integrated
in standard culture dishes (718,723). This technique allows a
more detailed analysis of the interaction between thyroid
hormone and contractile activity on muscle gene expression
(718).
[J.5] Skeleton
Background. Bone strength in adults is determined by
the acquisition of bone mass and mineral during growth and
the subsequent rate of bone loss throughout adulthood.
Normal euthy roid stat us is es sential for skeletal develop-
ment, growth, and mineralization a nd is a key determinant
of peak bone mass. T
3
also regulates bone remodeling in
adults, and thyroid status is an important determinant of the
rate of bone loss (724). Hypothyroidism results in a state of
red uced bone turnover, which if prolonged may result in to
accumulation of bone mass and mineral. By contrast, thy-
rotoxicosis accelerates bone turnover and loss resulting in
osteoporosis.
&
RECOMMENDATION 66
Thyroid hormones effects on postnatal linear growth and
bone maturation are best determined ex vivo by using
samples from animals harvested at defined ages.
Commentary. Mice should be genotyped and identified
at young postnatal ages while preserving their tail so that
linear growth velocity can be determined by weekly mea-
surement of tail length between postnatal days P7 to P70.
Maximum growth velocity occurs between P14 and P42 in
mice, and comparison of growth curves provides an accurate
indication of developmental delay. Neonatal P1 mice and
limbs from animals aged between P21 and P42 should be
stained with alizarin red (bone) and alcian blue (cartilage)
to investigate intrauterine skeletal development and the
progress of postnatal bone formation. Measurement of skull
dimensions, fontanelles, and suture areas provide an assess-
ment of intramembranous ossification; measures of bone
length and growth plate histomorphometry are accurate and
sensitive indicators of the progression of endochondral ossi-
fication (725–731).
&
RECOMMENDATION 67
The structural and mineralization response of adult bone to
alterations in thyroid status or disruption of thyroid hor-
mone action can be assessed using complementary imaging
techniques at different levels of resolution.
Commentary. Analysis of the skeleton in mi ce requi res
collection o f bone samples ex vivo.Itisnotpossibletode-
termine structural parameters accurately in vivo because
imaging modalities do not have sufficient resolution to dis-
criminate features of trabecular bone accurately and much
more informati on can be obtai ned from ex vivo samples using
complementary approaches and methods (Fig. 19). Bone
mineral content, bone length, and cortical bone measure-
ments are determined in long bones and vertebrae from
adults aged 12 weeks onwards by Faxitron digital point
projection x-ra y microradiography ( 725). Bone micro-
architecture and volumetric parameters require sophisti-
cated imaging and can be obtained by high resolution micro-
CT (to a maximum nominal resolution of 0.5–2 lm). Cortical
and trabecular bone volumes as proportions of tissue vol-
ume, trabecular thickness, trabecular separation, and tra-
becular number can be determined. Such parameters in the
mouse are at the limits of spatial resolution by micro-CT.
Detailed analysis of bone microarchitecture and micro-
mineralization density can be analyzed using specialist back
scattered electron-scanning electron microscopy (BSE-S EM)
techni ques (725–727,729–731). The effects of altered thyroid
status on linear growth and bone structural para meters are
significant and readily determined using these methods, al-
though specialist e quipment and data analysis methods are
required.
&
RECOMMENDATION 68
The thyroid hormone–induced changes in bone structure
and mineralization are reflected by functional abnormali-
ties, which can be investigated by determination of bio-
mechanical properties under loading.
Commentary. Thestrengthofboneisdeterminedusing
appropriate mechanical loading equipment (Fig. 20). Destruc-
tive three-point bend tests are performed on long bones from
adult mice to determine cortical bone biomechanical variables
from load displacement curves, including stiffness, maximum
load, fracture load, and energies dissipated at maximum load
and fracture. Trabecular bone compression strength can be
determined similarly in vertebrae. Cross-sectional area mea-
surements are obtained from Faxitron and micro-CT data and
used with biomechanical data to calculate ultimate stress, yield
stress, and modulus parameters (725). More detailed and spe-
cialist studies of material properties of bone at various resolu-
tion scales can be obtained using highly sophisticated
techniques and equipment in collaboration with engineering
and other specialist laboratories (732–734). Such measurements
provide accurate determination of bone strength together with
assessment of its material properties (brittleness or ductility)
that are affected significantly, for example, by loss of bone mass
and mineral content in thyrotoxicosis.
&
RECOMMENDATION 69a
Metabolic responses and changes in bone resorption and
formation in response to altered thyroid status can be as-
sessed indirectly by measurement of serum parameters.
Commentary. Biochemical analysis of mineral metabo-
lism and bone turno ver should include measurement of
serum Ca
2 +
,Mg
2 +
,PO
4
3 -
, and alkaline phosphatase,
which can be determined using standard auto-analyzers. In
addition to determination of thyroid status, other hormones
affecting bone can be assayed using commercial assays,
including corticostero ne, vitamin D, parathyroi d hormone,
and insulin-like growth factor-1. In addition, sex hormone
and gonadotropin status can be determined if required
(735). Indirect analysis of bone turnover can be obtained by
measurement of bone resorption a nd formation markers
using commercially available kits ( 725). C-terminal cross-
linked telopeptide of type I collagen a nd tartrate-resist ant
acid phosphatase (TRAcP) 5b are measured by ELISA and
solid-phase immuno-fixed enzyme activity assay as
140 BIANCO ET AL.
indicators of bone resorptio n. N- terminal p rope ptide of
type 1 procollagen and osteocalcin are also measured b y
ELISA and IRMA as m arkers of bone formation. Assess-
ment of bone turnover markers should ideally be per-
formed ideally in the morning o n fasted animals to mitigate
the c onfounding diurnal and dietary effects on bone turn-
over. Older rodents have low levels of bone turnover in the
basal state and determination cannot be performed reliably
after about 12–16 weeks of age as levels of markers will be
close to the limits of detection. It is important to note there
are sex- and strain-specific differences in bone mass and
turnover and care in experimental design should be taken
to ensure littermates are compared where poss ible. Never-
theless, measurement of bone turnover using sur rogate
bonemarkersisanestablishedmethodinthefieldthat
provides additional information to support imaging and
histomorphometry data.
&
RECOMMENDATION 69b
The effects of thyroid hormone on bone turnover might in-
volve effects on osteoclastic bone resorption, osteoblastic bone
formation, or a combination of both. Biochemical analysis of
markers of bone turnover in response to thyroid hormone
provides indirect crude assessment of these parameters, but
this must also be investigated directly in the bone tissue.
Commentary. Hypothyroidism results in reduced bone
turnover affecting both bone formation and resorption.
FIG. 20. Instron5534 load frame apparatus for destructive
three-point bend testing of bone strength. Custom mounts
incorporate rounded supports and loading pins that mini-
mize cutting and shear forces, thus enabling biomechanical
parameters of bone strength to be determined accurately.
Courtesy of Dr. Graham Williams.
FIG. 19. Scheme for collection of bone samples, tissue fixation, and methods of analysis. The usage of this chart can
maximize phenotype information obtained from juvenile and adult mice. qBSE-SEM, quantitative back-scattered electron-
scanning electron microscopy. Courtesy of Dr. Graham Williams.
INVESTIGATING THYROID HORMONE ECONOMY AND ACTION 141
Thyrotoxicosis accelerates both parameters with a greater
effect on bone resorption resulting in bone loss. To inves-
tigate and quantify the cellular basis for altered bone
turnover, dynamic measurem ents by histomorphometry
are required. Mice are given two intraperitoneal injections
of the fluorochrome calcein (15 mg/kg in saline with 2%
sodium bicarbonate) approximately 7 and 4 days prior to
sacrifice, alt hough the precise interval can be varied and
should be determined empirically depending on the locally
available analysis techniques. Dynamic bone histomor-
phometry parameters should be determined according to
the sta ndard system of nomenclature (736). Bone resorp -
tion is investigated in bone sections stained for TRAcP,
and osteoclast numbers and surface determined by light
microscopy and normalized to total bone surface. Cortical
and trabecular bone resorption surfaces in vivo can also
be determined using specialized BSE-SEM techniques
(725,727,729), although suc h methods a re not routinely or
widely available. Bone formation should be quantified
in dual calcein-labeled bone by standard histomorpho-
metry and indices of bone formation determined using
ImageJ (http://rsbweb.nih.gov/ij) according to the Amer-
ican Society for Bone and Mineral Research standard sys-
tem o f nomenclature (736 ). Alternative high resolution
methods using confocal microscopy can be performed in
specialist laboratories (725,737). Gene expression studies
can be performed in paraffin-embedded decalcified sec-
tions of bone and cartilage (in situ hybridization and IHC),
using RNA extracted from tissue samples (mRNA expres-
sion and profiling) and in skeletal tissue extracts (protein
expression and enzyme activity assays) (726,727,729 –
731,738–740).
&
RECOMMENDATION 70a
The mechanisms of thyroid hormone action in cartilage can
be studied in chondrocyte cultures prepared from limbs or
rib cages of neonatal mice.
Commentary. Bone is a complex organ and the cellular
responses to thyroid hormone can be determined by in vivo
and tissue studies. A deeper understanding of the molec-
ular mechanisms of T
3
action and t he signaling pathways
involved in mediating T
3
responses in bone cell lineages can
be obtained by studying thyroid hormone effects in pri-
mary cell cultures of sk eletal cells. Primary chondrocytes
are prepared from neonatal mice (741,742) and prolifera-
tion, differentiation (collagen X mRNA expression, alkaline
phosphatase a ctivity, alcian blue staining), apoptosis, and
gene expression responses (Ihh, PTHrP, Sox9 by RT-qPCR)
to T
3
treatment determined after 7, 14, and 28 days. It
should be noted that chondrocytes are difficult to study
in vitro because they tend to de-differentiate and acquire a
fibroblast-like phenotype in monolayer culture after a few
days.Thisdoesnotposeaproblemforgeneexpression
studie s if investigation of acute responses to T
3
treatment is
limited to analysis of freshly cultured cells within the first
48 hours. However, in studies to investigate effects of thy-
roid hormone on chondrogenesis and cell proliferation over
longer periods of time as already indicated, 3D cultures
should be performed in agarose suspension to prevent the
tenden cy for de-differe ntiation of chondrocytes when cul-
tured for prolonged perio ds in mono layers (653).
&
RECOMMENDATION 70b
Osteoblast cultures can be employed to investigate the
mechanisms of thyroid hormone action on cell differenti-
ation, as well as on functional bone mineralization assays
in vitro. Such cultures are usually prepared by collagenase
digestion of calvariae obtained from neonatal mice using
standard techniques.
Commentary. Osteoblasts express TRs and, like chon-
drocytes, respond directly to the actions of T
3
. Primary
calvarial osteoblasts are prepared from P4 pups and prolif-
eration, differentiation (osteocalcin and osterix mRNA ex-
pression, alkaline phosphatase activity, alizarin red and von
Kossa mineralization assays), and gene expression (Runx-2,
OPG, receptor activator of nuclear factor kappa B-ligand
[RANKL] by RT-qPCR) responses are usually determined
after 7, 14, and 28 days (738–740,743).
&
RECOMMENDATION 70c
The effects of thyroid hormone on osteoclastogenesis and
function can be determined in osteoclast cultures prepared
from 6-week-old mouse long bones. The protocols used
include treating extracted bone marrow cells with mono-
cyte colony stimulating factor and RANKL to induce os-
teoclast formation in vitro.
Commentary. Bone marrow is isolated from long bones of
P42 mice and primary cultures treated with monocyte colony
stimulating factor and RANKL to induce osteoclast formation
in vitro (725,729,740). Osteoclast numbers (following TRAP
staining) and cell differentiation (TRAP activity, calcitonin re-
ceptor, cathepsin K mRNA expression) are determined after
days 0, 3, 6, and 9 days. Osteoclast activity can also be deter-
mined by dentine resorption assay using standard techniques
employed previously in studies of T
3
action (725,729). The
mechanism of thyroid hormone action on osteoclast function is
currently unclear, and it is uncertain whether the major re-
sponse to increase osteoclastic bone resorption represents a
direct effect of T
3
in osteoclasts or whether the response is a
secondary effect mediated via the actions of T
3
in osteoblasts,
bone marrow stromal cells, or other factors.
ACKNOWLEDGMENTS
ThetaskforcewishestothankMs.BobbiSmith,Executive
Director, ATA, and Ms. Adonia Coates and Ms. Sharleene Cano,
ATA staff members, for their assistance to the task force, their
expert help, and their support; Mr. Cezar Bianchi for valuable
assistance with illustrations; and Ms. Yanette Pericich and Cris-
tina Andrade (Division of Endocrinology, University of Miami
Miller School of Medicine) for their assistance preparing the text.
The task force also wishes to thank the following colleagues who
provided insightful comments and/or support materials:
Sherine Abdalla, MD (Division of Endocrinology, Uni-
versity of Miami Miller School of Medicine)
Anita Boelen, PhD (Department of Endocrinology, Aca-
demic Medical Center, Amsterdam, The Netherlands)
Denise P. Carvalho, MD, PhD (Carlos Chagas Filho In-
stitute of Biophysics, Federal University of Rio de Ja-
neiro, Rio de Janeiro, Brazil)
Orsolya Doha
´
n, MD, PhD (Department of Clinical Studies,
Faculty of Health Science; First Department of Internal
142 BIANCO ET AL.
Medicine, Faculty of Medicicne; Semmelweis Medical
School, Budapest, Hungary)
James Fagin, MD (Division of Endocrinology, Memorial
Sloan-Kettering Cancer Center, New York, NY)
Tatiana Fonseca, PhD (Division of Endocrinology, Uni-
versity of Miami Miller School of Medicine)
Ronald Koenig, MD, PhD (Division of Metabolism, En-
docrinology, and Diabetes, University of Michigan, Ann
Arbor, MI)
Vania Nose, MD, PhD (Department of Pathology, Massa-
chusetts General Hospital, Boston, MA)
Miriam O. Ribeiro, PhD (Department of Biological Sciences
and Health Science Center, MacKenzie Presbiterian
University, Sa
˜
o Paulo, Brazil)
Thomas Scanlan, PhD (Department of Physiology and
Pharmacology, School of Medicine, Oregon Health and
Science University, Portland, OR)
Steven J. Soldin, PhD (Department of Laboratory Medi-
cine, National Institutes of Medicine, Washington, DC)
Cintia Ueta, PhD (Department of Anatomy, University of
Sa
˜
o Paulo, Sa
˜
o Paulo, Brazil)
This work was supported in part by the intramural research
program of NIDDK at the National Institutes of Health (to D.F.),
NIDDK grants 2R01-DK58538-08, 1R01-DK065055-02, R01-
DK77148 (to A.C.B.), R37DK015070 (to X.H.L. and S.R.) and R21
HD 072526-02 (to V.A.G.), NICHD grant R03 HD061901 (to
P.A.K.), SAF2012-32 491 from MINECO and S2010/BMD-2423
from CAM (to M.J.O.), U.S.–Israel Binational Science Founda-
tion (to G.A.), National Science Foundation of Hungary (OTKA
K81226) and the European Community’s Seventh Framework
Programme (FP7/2007–2013) under grant agreement no.
259772 (to B.G.), and the ZonMw Clinical Fellowship Grant
(90700412) and a ZonMW TOP Grant (91212044) (to R.P.P.).
DISCLOSURE STATEMENT
Disclosure Information for 2 years before December 2012
and the known future as of November 2013: all authors de-
clare no significant financial conflicts of interest to disclose.
REFERENCES
1. Toda S, Aoki S, Uchihashi K, Matsunobu A, Yamamoto M,
Ootani A, Yamasaki F, Koike E, Sugihara H 2011 Culture
models for studying thyroid biology and disorders. ISRN
Endocrinol 2011:275782.
2. Rotman-Pikielny P, Hirschberg K, Maruvada P, Suzuki K,
Royaux IE, Green ED, Kohn LD, Lippincott-Schwartz J, Yen
PM 2002 Retention of pendrin in the endoplasmic reticu-
lum is a major mechanism for Pendred syndrome. Hum
Mol Genet 11:2625–2633.
3. Taylor JP, Metcalfe RA, Watson PF, Weetman AP, Trembath
RC 2002 Mutations of the PDS gene, encoding pendrin, are
associated with protein mislocalization and loss of iodide
efflux: implications for thyroid dysfunction in Pendred
syndrome. J Clin Endocrinol Metab 87:1778–1784.
4. Weiss SJ, Philp NJ, Ambesi-Impiombato FS, Grollman EF
1984 Thyrotropin-stimulated iodide transport mediated by
adenosine 3¢,5¢-monophosphate and dependent on protein
synthesis. Endocrinology 114:1099–1107.
5. Weiss SJ, Philp NJ, Grollman EF 1984 Effect of thyrotropin
on iodide efflux in FRTL-5 cells mediated by Ca
2 +
. En-
docrinology 114:1108–1113.
6. Weiss SJ, Philp NJ, Grollman EF 1984 Iodide transport in a
continuous line of cultured cells from rat thyroid. En-
docrinology 114:1090–1098.
7. Doha
´
n O, Portulano C, Basquin C, Reyna-Neyra A, Amzel
LM, Carrasco N 2007 The Na + /I symporter (NIS) mediates
electroneutral active transport of the environmental pollut-
ant perchlorate. Proc Natl Acad Sci USA 104:20250–20255.
8. Biddinger P 2009 Normal anatomy and histology. In: Ni-
kiforov Y, Thompson L, Biddinger P (eds.) Diagnostic
Surgical Pathology and Molecular Genetics of the Thyroid.
Lippincott Williams and Wilkins, Philadelphia, pp 1–10.
9. Penel C, Rognoni JB, Bastiani P 1987 Thyroid autoregula-
tion: impact on thyroid structure and function in rats. Am J
Physiol Endocrinol Metab 253:E165–E172.
10. Fujita H 1988 Functional morphology of the thyroid. Int
Rev Cytol 113:145–185.
11. Delverdier M, Cabanie P, Roome N, Enjalbert F, Van Ha-
verbeke G 1991 Critical analysis of the histomorphometry
of rat thyroid after treatment with thyroxin and pro-
pylthiouracil. Ann Rech Vet 22:373–378.
12. Leatherland JF, Sonstegard RA 1980 Structure of thyroid
and adrenal glands in rats fed diets of Great Lakes coho
salmon (Oncorhynchus kisutch). Environ Res 23:77–86.
13. Kmiec Z, Kotlarz G, Smiechowska B, Mysliwski A 1998 The
effect of fasting and refeeding on thyroid follicule structure
and thyroid hormone levels in young and old rats. Arch
Gerontol Geriatr 26:161–175.
14. Denef JF, Cordier AC, Mesquita M, Haumont S 1979 The
influence of fixation procedure, embedding medium and
section thickness on morphometric data in thyroid gland.
Histochemistry 63:163–171.
15. Asmis LM, Kaempf J, Von Gruenigen C, Kimura ET,
Wagner HE, Studer H 1996 Acquired and naturally oc-
curring resistance of thyroid follicular cells to the growth
inhibitory action of transforming growth factor-beta 1
(TGF-beta 1). J Endocrinol 149:485–496.
16. Cauvi D, Penel C, Nlend MC, Venot N, Allasia C, Chabaud
O 2000 Regulation of thyroid cell volumes and fluid
transport: opposite effects of TSH and iodide on cultured
cells. Am J Physiol Endocrinol Metab 279:E546–E553.
17. Schurch M, Peter HJ, Gerber H, Studer H 1990 Cold follicles
in a multinodular human goiter arise partly from a failing
iodide pump and partly from deficient iodine organifica-
tion. J Clin Endocrinol Metab 71:1224–1229.
18. Gerber H, Studer H, and von Grunigen C 1985 Paradoxical
effects of thyrotropin on diffusion of thyroglobulin in the
colloid of rat thyroid follicles after long term thyroxine
treatment. Endocrinology 116:303–310.
19. Peter HJ, Gerber H, Studer H, Becker DV, Peterson ME
1987 Autonomy of growth and of iodine metabolism in
hyperthyroid feline goiters transplanted onto nude mice.
J Clin Invest 80:491–498.
20. Senou M, Costa MJ, Massart C, Thimmesch M, Khalifa C,
Poncin S, Boucquey M, Gerard AC, Audinot JN, Dessy C,
Ruf J, Feron O, Devuyst O, Guiot Y, Dumont JE, Van Sande
J, Many MC 2009 Role of caveolin-1 in thyroid phenotype,
cell homeostasis, and hormone synthesis: in vivo study of
caveolin-1 knockout mice. Am J Physiol Endocrinol Metab
297:E438–E451.
21. Elbast M, Wu TD, Guiraud-Vitaux F, Petiet A, Hindie E,
Champion C, Croisy A, Guerquin-Kern JL, Colas-Linhart N
2008 [Kinetics of intracolloidal iodine within the thyroid of
newborn rats. Direct imagery using secondary ion mass
spectrometry]. C R Biol 331:13–22. (In French.)
INVESTIGATING THYROID HORMONE ECONOMY AND ACTION 143
22. Gerard AC, Poncin S, Caetano B, Sonveaux P, Audinot JN,
Feron O, Colin IM, Soncin F 2008 Iodine deficiency induces
a thyroid stimulating hormone-independent early phase
of microvascular reshaping in the thyroid. Am J Pathol
172:748–760.
23. Burgi-Saville ME, Gerber H, Peter HJ, Paulsson M, Aes-
chlimann D, Glaser C, Kaempf J, Ruchti C, Sidiropoulos I,
Burgi U 1997 Expression patterns of extracellular matrix
components in native and cultured normal human thyroid
tissue and in human toxic adenoma tissue. Thyroid 7:
347–356.
24. Kimura ET, Kopp P, Zbaeren J, Asmis LM, Ruchti C, Maciel
RM, Studer H 1999 Expression of transforming growth
factor beta1, beta2, and beta3 in multinodular goiters and
differentiated thyroid carcinomas: a comparative study.
Thyroid 9:119–125.
25. Wapnir IL, van de Rijn M, Nowels K, Amenta PS, Walton
K, Montgomery K, Greco RS, Dohan O, Carrasco N 2003
Immunohistochemical profile of the sodium/iodide sym-
porter in thyroid, breast, and other carcinomas using high
density tissue microarrays and conventional sections. J Clin
Endocrinol Metab 88:1880–1888.
26. Chen CR, Chazenbalk GD, Wawrowsky KA, McLachlan
SM, Rapoport B 2006 Evidence that human thyroid cells
express uncleaved, single-chain thyrotropin receptors on
their surface. Endocrinology 147:3107–3113.
27. Bernier-Valentin F, Kostrouch Z, Rabilloud R, Munari-
Silem Y, Rousset B 1990 Coated vesicles from thyroid cells
carry iodinated thyroglobulin molecules. First indication
for an internalization of the thyroid prohormone via a
mechanism of receptor-mediated endocytosis. J Biol Chem
265:17373–17380.
28. Kopp P 2005 Thyroid hormone synthesis: thyroid iodine
metabolism. In: Braverman L, Utiger R (eds.) Werner and
Ingbar’s The Thyroid: A Fundamental and Clinical Text,
9th edition. Lippincott, Williams & Wilkins, Philadelphia,
pp 52–76.
29. Dai G, Levy O, Carrasco N 1996 Cloning and characteriza-
tion of the thyroid iodide transporter. Nature 379:458–460.
30. Doha
´
n O, De la Vieja A, Paroder V, Riedel C, Artani M,
Reed M, Ginter CS, Carrasco N 2003 The sodium/iodide
symporter (NIS): characterization, regulation, and medical
significance. Endocr Rev 24:48–77.
31. Nilsson M, Bjo
¨
rkman U, Ekholm R, Ericson LE 1990 Iodide
transport in primary cultured thyroid follicle cells: evidence
of a TSH-regulated channel mediating iodide efflux selec-
tively across the apical domain of the plasma membrane.
Eur J Cell Biol 52:270–281.
32. Golstein P, Abramow M, Dumont JE, Beauwens R 1992 The
iodide channel of the thyroid: a plasma membrane vesicle
study. Am J Physiol 263:C590–C597.
33. Pesce L, Bizhanova A, Caraballo JC, Westphal W, Butti ML,
Comellas A, Kopp P 2012 TSH regulates pendrin mem-
brane abundance and enhances iodide efflux in thyroid
cells. Endocrinology 153:512–521.
34. Shcheynikov N, Yang D, Wang Y, Zeng W, Karniski LP, So I,
Wall SM, Muallem S 2008 The Slc26a4 transporter functions
as an electroneutral Cl-/I-/HCO3- exchanger: role of Slc26a4
and Slc26a6 in I- and HCO3- secretion and in regulation of
CFTR in the parotid duct. J Physiol 586:3813–3824.
35. Dossena S, Bizhanova A, Nofziger C, Bernardinelli E,
Ramsauer J, Kopp P, Paulmichl M 2011 Identification of
allelic variants of pendrin (SLC26A4) with loss and gain of
function. Cell Physiol Biochem 28:467–476.
36. Dossena S, Rodighiero S, Vezzoli V, Bazzini C, Sironi C,
Meyer G, Furst J, Ritter M, Garavaglia ML, Fugazzola L,
Persani L, Zorowka P, Storelli C, Beck-Peccoz P, Botta
´
G,
Paulmichl M 2006 Fast fluorometric method for measuring
pendrin (SLC26A4) Cl-/I- transport activity. Cell Physiol
Biochem 18:67–74.
37. Dossena S, Rodighiero S, Vezzoli V, Nofziger C, Salvioni E,
Boccazzi M, Grabmayer E, Botta G, Meyer G, Fugazzola L,
Beck-Peccoz P, Paulmichl M 2009 Functional characteriza-
tion of wild-type and mutated pendrin (SLC26A4), the
anion transporter involved in Pendred syndrome. J Mol
Endocrinol 43:93–103.
38. Tran N, Valentin-Blasini L, Blount BC, McCuistion CG,
Fenton MS, Gin E, Salem A, Hershman JM 2008 Thyroid-
stimulating hormone increases active transport of per-
chlorate into thyroid cells. Am J Physiol Endocrinol Metab
294:E802–E806.
39. Nilsson M, Bjo
¨
rkman U, Ekholm R, Ericson LE 1992 Po-
larized efflux of iodide in porcine thyrocytes occurs via a
cAMP-regulated iodide channel in the apical plasma
membrane. Acta Endocrinol (Copenh) 126:67–74.
40. Gillam MP, Sidhaye A, Lee EJ, Rutishauser J, Waeber
Stephan C, Kopp P 2004 Functional characterization of
pendrin in a polarized cell system: evidence for pendrin-
mediated apical iodide efflux. J Biol Chem 279:13004–
13010.
41. Zuckier LS, Dohan O, Li Y, Chang CJ, Carrasco N, Da-
dachova E 2004 Kinetics of perrhenate uptake and com-
parative biodistribution of perrhenate, pertechnetate, and
iodide by NaI symporter-expressing tissues in vivo. J Nucl
Med 45:500–507.
42. Reed-Tsur MD, De la Vieja A, Ginter CS, Carrasco N 2008
Molecular characterization of V59E NIS, a Na + /I- sym-
porter mutant that causes congenital I- transport defect.
Endocrinology 149:3077–3084.
43. Scholz IV, Cengic N, Baker CH, Harrington KJ, Maletz K,
Bergert ER, Vile R, Goke B, Morris JC, Spitzweg C 2005
Radioiodine therapy of colon cancer following tissue-
specific sodium iodide symporter gene transfer. Gene Ther
12:272–280.
44. Spitzweg C, Morris JC 2002 The sodium iodide symporter:
its pathophysiological and therapeutic implications. Clin-
ical Endocrinol (Oxf) 57:559–574.
45. Grasberger H, De Deken X, Mayo OB, Raad H, Weiss M,
Liao XH, Refetoff S 2012 Mice deficient in dual oxidase
maturation factors are severely hypothyroid. Mol Endo-
crinol 26:481–492.
46. Rocchi R, Kunavisarut T, Ladenson P, Caturegli P 2006
Thyroid uptake of radioactive iodine and scintigraphy in
mice. Thyroid 16:705–706.
47. van den Hove MF, Croizet-Berger K, Jouret F, Guggino SE,
Guggino WB, Devuyst O, Courtoy PJ 2006 The loss of the
chloride channel, ClC-5, delays apical iodide efflux and
induces a euthyroid goiter in the mouse thyroid gland.
Endocrinology 147:1287–1296.
48. Reinfelder J, Maschauer S, Foss CA, Nimmagadda S, Fre-
mont V, Wolf V, Weintraub BD, Pomper MG, Szkudlinski
MW, Kuwert T, Prante O. Effects of recombinant human
thyroid-stimulating hormone superagonists on thyroidal
uptake of 18F-fluorodeoxyglucose and radioiodide. Thy-
roid 21:783–792.
49. Colzani RM, Alex S, Fang SL, Braverman LE, Emerson CH
1998 The effect of recombinant human thyrotropin (rhTSH)
on thyroid function in mice and rats. Thyroid 8:797–801.
144 BIANCO ET AL.
50. Hilditch TE, Horton PW, McCruden DC, Young RE,
Alexander WD 1982 Defects in intrathyroid binding of io-
dine and the perchlorate discharge test. Acta Endocrinol
(Copenh) 100:237–244.
51. Atterwill CK, Collins P, Brown CG, Harland RF 1987 The
perchlorate discharge test for examining thyroid function in
rats. J Pharmacol Methods 18:199–203.
52. Coelho-Palermo Cunha G, van Ravenzwaay B 2007 Stan-
dardization of the perchlorate discharge assay for thyroid
toxicity testing in rats. Regul Toxicol Pharmacol 48:270–
278.
53. Dingli D, Kemp BJ, O’Connor MK, Morris JC, Russell SJ, Lowe
VJ 2006 Combined I-124 positron emission tomography/
computed tomography imaging of NIS gene expression in
animal models of stably transfected and intravenously
transfected tumor. Mol Imaging Biol 8:16–23.
54. Unger J, Boeynaems JM, Van Herle A, Van Sande J, Roc-
mans P, Mockel J 1979 In vitro nonbutanol-extractable io-
dine release in dog thyroid. Endocrinology 105:225–231.
55. Franken PR, Guglielmi J, Vanhove C, Koulibaly M, Defrise
M, Darcourt J, Pourcher T 2010 Distribution and dynamics
of (99m)Tc-pertechnetate uptake in the thyroid and other
organs assessed by single-photon emission computed to-
mography in living mice. Thyroid 20:519–526.
56. Brandt MP, Kloos RT, Shen DH, Zhang X, Liu YY, Jhiang
SM 2012 Micro-single-photon emission computed tomog-
raphy image acquisition and quantification of sodium-
iodide symporter-mediated radionuclide accumulation in
mouse thyroid and salivary glands. Thyroid 22:617–624.
57. Emanuelsson M 2006 Development of an animal in vivo
124I-microPET/microCAT imaging model of the thyroid
[MS thesis]. Medical Radiation Physics Clinical Sciences.
Lund University, Lund, Sweden.
58. Chakravarty D, Santos E, Ryder M, Knauf JA, Liao XH,
West BL, Bollag G, Kolesnick R, Thin TH, Rosen N, Zan-
zonico P, Larson SM, Refetoff S, Ghossein R, Fagin JA 2011
Small-molecule MAPK inhibitors restore radioiodine in-
corporation in mouse thyroid cancers with conditional
BRAF activation. J Clin Invest 121:4700–4711.
59. Renault G, Bonnin P, Marchiol-Fournigault C, Gregoire JM,
Serriere S, Richard B, Fradelizi D 2006 [High-resolution
ultrasound imaging of the mouse]. J Radiol 87:1937–1945.
(In French.)
60. Foster FS, Zhang MY, Zhou YQ, Liu G, Mehi J, Cherin E,
Harasiewicz KA, Starkoski BG, Zan L, Knapik DA,
Adamson SL 2002 A new ultrasound instrument for in vivo
microimaging of mice. Ultrasound Med Biol 28:1165–1172.
61. Mancini M, Vergara E, Salvatore G, Greco A, Troncone G,
Affuso A, Liuzzi R, Salerno P, Scotto di Santolo M, Santoro
M, Brunetti A, Salvatore M 2009 Morphological ultrasound
microimaging of thyroid in living mice. Endocrinology
150:4810–4815.
62. Wassen FW, Schiel AE, Kuiper GG, Kaptein E, Bakker O,
Visser TJ, Simonides WS 2002 Induction of thyroid
hormone-degrading deiodinase in cardiac hypertrophy and
failure. Endocrinology 143:2812–2815.
63. Simonides WS, Mulcahey MA, Redout EM, Muller A,
Zuidwijk MJ, Visser TJ, Wassen FW, Crescenzi A, da-Silva
WS, Harney J, Engel FB, Obregon MJ, Larsen PR, Bianco
AC, Huang SA 2008 Hypoxia-inducible factor induces local
thyroid hormone inactivation during hypoxic-ischemic
disease in rats. J Clin Invest 118:975–983.
64. Ueta CB, Oskouei BN, Olivares EL, Pinto JR, Correa MM,
Simovic G, Simonides WS, Hare JM, Bianco AC 2012 Ab-
sence of myocardial thyroid hormone inactivating deiodi-
nase results in restrictive cardiomyopathy in mice. Mol
Endocrinol 26:809–818.
65. Jo S, Kallo
´
I, Bardoczi Z, Arrojo EDR, Zeold A, Liposits Z,
Oliva A, Lemmon VP, Bixby JL, Gereben B, Bianco AC 2012
Neuronal hypoxia induces hsp40-mediated nuclear import
of type 3 deiodinase as an adaptive mechanism to reduce
cellular metabolism. J Neurosci
32:8491–8500.
66. Bianco AC, Silva JE 1988 Cold exposure rapidly induces
virtual saturation of brown adipose tissue nuclear T
3
re-
ceptors. Am J Physiol 255:E496–E503.
67. Ferrara AM, Liao XH, Gil-Ibanez P, Marcinkowski T, Ber-
nal J, Weiss RE, Dumitrescu AM, Refetoff S 2013 Changes
in thyroid status during perinatal development of MCT8-
deficient male mice. Endocrinology 154:2533–2541.
68. Pohlenz J, Maqueem A, Cua K, Weiss RE, Van Sande J,
Refetoff S 1999 Improved radioimmunoassay for mea-
surement of mouse thyrotropin in serum: strain differences
in thyrotropin concentration and thyrotroph sensitivity to
thyroid hormone. Thyroid 9:1265–1271.
69. Hall JA, Ribich S, Christoffolete MA, Simovic G, Correa-
Medina M, Patti ME, Bianco AC 2010 Absence of thyroid
hormone activation during development underlies a per-
manent defect in adaptive thermogenesis. Endocrinology
151:4573–4582.
70. Calvo R, Morreale de Escobar G, Escobar del Rey F, Obregon
MJ 1997 Maternal nonthyroidal illness and fetal thyroid
hormone status, as studied in the streptozotocin-induced
diabetes mellitus rat model. Endocrinology 138:1159–1169.
71. Pedraza PE, Obregon MJ, Escobar-Morreale HF, Escobar del
Rey F, Morreale de Escobar G 2006 Mechanisms of adapta-
tion to iodine deficiency in rats: thyroid status is tissue spe-
cific. Its relevance for man. Endocrinology 147:2098–2108.
72. Zavacki AM, Ying H, Christoffolete MA, Aerts G, So E,
Harney JW, Cheng SY, Larsen PR, Bianco AC 2005 Type 1
iodothyronine deiodinase is a sensitive marker of peripheral
thyroid status in the mouse. Endocrinology 146:1568–1575.
73. Soukhova N, Soldin OP, Soldin SJ 2004 Isotope dilution
tandem mass spectrometric method for T4/T3. Clin Chim
Acta 343:185–190.
74. Piehl S, Heberer T, Balizs G, Scanlan TS, Kohrle J 2008
Development of a validated liquid chromatography/
tandem mass spectrometry method for the distinction of
thyronine and thyronamine constitutional isomers and for
the identification of new deiodinase substrates. Rapid
Commun Mass Spectrom 22:3286–3296.
75. Go
¨
the S, Wang Z, Ng L, Kindblom JM, Barros AC, Ohlsson
C, Vennstrom B, Forrest D 1999 Mice devoid of all known
thyroid hormone receptors are viable but exhibit disorders
of the pituitary-thyroid axis, growth, and bone maturation.
Genes Dev 13:1329–1341.
76. Perret J, Ludgate M, Libert F, Gerard C, Dumont JE, Vassart
G, Parmentier M 1990 Stable expression of the human TSH
receptor in CHO cells and characterization of differentially
expressing clones. Biochem Biophys Res Commun 171:
1044–1050.
77. Brooker G, Harper JF, Terasaki WL, Moylan RD 1979
Radioimmunoassay of cyclic AMP and cyclic GMP. Adv
Cyclic Nucleotide Res 10:1–33.
78. Persani L, Asteria C, Tonacchera M, Vitti P, Krishna V,
Chatterjee K, Beck-Peccoz P 1994 Evidence for the secretion
of thyrotropin with enhanced bioactivity in syndromes of
thyroid hormone resistance. J Clin Endocrinol Metab
78:1034–1039.
INVESTIGATING THYROID HORMONE ECONOMY AND ACTION 145
79. Moeller LC, Alonso M, Liao X, Broach V, Dumitrescu A,
Van Sande J, Montanelli L, Skjei S, Goodwin C, Grasberger
H, Refetoff S, Weiss RE 2007 Pituitary-thyroid setpoint
and thyrotropin receptor expression in consomic rats. En-
docrinology 148:4727–4733.
80. Yamada M, Saga Y, Shibusawa N, Hirato J, Murakami M,
Iwasaki T, Hashimoto K, Satoh T, Wakabayashi K, Taketo
MM, Mori M 1997 Tertiary hypothyroidism and hypergly-
cemia in mice with targeted disruption of the thyrotropin-
releasing hormone gene. Proc Natl Acad Sci USA 94:
10862–10867.
81. Moeller LC, Kimura S, Kusakabe T, Liao XH, Van Sande J,
Refetoff S 2003 Hypothyroidism in thyroid transcription
factor 1 haploinsufficiency is caused by reduced expression
of the thyroid-stimulating hormone receptor. Mol En-
docrinol 17:2295–2302.
82. Astapova I, Vella KR, Ramadoss P, Holtz KA, Rodwin BA,
Liao XH, Weiss RE, Rosenberg MA, Rosenzweig A, Hol-
lenberg AN 2011 The nuclear receptor corepressor (NCoR)
controls thyroid hormone sensitivity and the set point of
the hypothalamic-pituitary-thyroid axis. Mol Endocrinol
25:212–224.
83. Morreale de Escobar G, Pastor R, Obregon MJ, Escobar del
Rey F 1985 Effects of maternal hypothyroidism on the
weight and thyroid hormone content of rat embryonic tis-
sues, before and after onset of fetal thyroid function. En-
docrinology 117:1890–1900.
84. Gereben B, Zavacki AM, Ribich S, Kim BW, Huang SA,
Simonides WS, Zeold A, Bianco AC 2008 Cellular and
molecular basis of deiodinase-regulated thyroid hormone
signaling. Endocr Rev 29:898–938.
85. Larsen PR, Silva JE, Kaplan MM 1981 Relationships between
circulating and intracellular thyroid hormones: physiological
and clinical implications. Endocr Rev 2:87–102.
86. Bianco AC, Silva JE 1987 Nuclear 3,5,3¢-triiodothyronine
(T3) in brown adipose tissue: receptor occupancy and
sources of T
3
as determined by in vivo techniques. En-
docrinology 120:55–62.
87. Silva JE, Dick TE, Larsen PR 1978 The contribution of local
tissue thyroxine monodeiodination to the nuclear 3,5,3¢-
triiodothyronine in pituitary, liver, and kidney of euthyroid
rats. Endocrinology 103:1196–1207.
88. Silva JE, Larsen PR 1978 Contributions of plasma triiodo-
thyronine and local thyroxine monodeiodination to triio-
dothyronine to nuclear triiodothyronine receptor saturation
in pituitary, liver, and kidney of hypothyroid rats. Further
evidence relating saturation of pituitary nuclear triiodo-
thyronine receptors and the acute inhibition of thyroid-
stimulating hormone release. J Clin Invest 61:1247–1259.
89. Crantz FR, Silva JE, Larsen PR 1982 Analysis of the sources
and quantity of 3,5,3¢-triiodothyronine specifically bound
to nuclear receptors in rat cerebral cortex and cerebellum.
Endocrinology 110:367–375.
90. Gouveia CH, Christoffolete MA, Zaitune CR, Dora JM,
Harney JW, Maia AL, Bianco AC 2005 Type 2 iodothyro-
nine selenodeiodinase is expressed throughout the mouse
skeleton and in the MC3T3-E1 mouse osteoblastic cell line
during differentiation. Endocrinology 146:195–200.
91. Salvatore D, Tu H, Harney JW, Larsen PR 1996 Type 2
iodothyronine deiodinase is highly expressed in human
thyroid. J Clin Invest 98:962–968.
92. Dentice M, Marsili A, Ambrosio R, Guardiola O, Sibilio A,
Paik JH, Minchiotti G, DePinho RA, Fenzi G, Larsen PR,
Salvatore D 2011 The FoxO3/type 2 deiodinase pathway is
required for normal mouse myogenesis and muscle re-
generation. J Clin Invest 120:4021–4030.
93. Curcio C, Baqui MM, Salvatore D, Rihn BH, Mohr S,
Harney JW, Larsen PR, Bianco AC 2001 The human type 2
iodothyronine deiodinase is a selenoprotein highly ex-
pressed in a mesothelioma cell line. J Biol Chem 276:30183–
30187.
94. Hosoi Y, Murakami M, Mizuma H, Ogiwara T, Imamura
M, Mori M 1999 Expression and regulation of type II io-
dothyronine deiodinase in cultured human skeletal muscle
cells. J Clin Endocrinol Metab 84:3293–3300.
95. Murakami M, Araki O, Morimura T, Hosoi Y, Mizuma H,
Yamada M, Kurihara H, Ishiuchi S, Tamura M, Sasaki T,
Mori M 2000 Expression of type II iodothyronine deiodinase
in brain tumors. J Clin Endocrinol Metab 85:4403–4406.
96. Murakami M, Araki O, Hosoi Y, Kamiya Y, Morimura T,
Ogiwara T, Mizuma H, Mori M 2001 Expression and reg-
ulation of type II iodothyronine deiodinase in human thy-
roid gland. Endocrinology 142:2961–2967.
97. Maeda A, Toyoda N, Yasuzawa-Amano S, Iwasaka T,
Nishikawa M 2003 Type 2 deiodinase expression is stim-
ulated by growth factors in human vascular smooth muscle
cells. Mol Cell Endocrinol 200:111–117.
98. Wajner SM, dos Santos Wagner M, Melo RC, Parreira GG,
Chiarini-Garcia H, Bianco AC, Fekete C, Sanchez E, Lechan
RM, Maia AL 2007 Type 2 iodothyronine deiodinase is
highly expressed in germ cells of adult rat testis. J En-
docrinol 194:47–54.
99. DiStefano JJ, Malone TK, Jang M 1982 Comprehensive ki-
netics of thyroxine distribution and metabolism in blood
and tissue pools of the rat from only six blood samples:
dominance of large, slowly exchanging tissue pools. En-
docrinology 111:108–117.
100. Nguyen TT, Chapa F, DiStefano JJ 3rd 1998 Direct mea-
surement of the contributions of type I and type II
5¢-deiodinases to whole body steady state 3,5,3¢-triiodo-
thyronine production from thyroxine in the rat. Endo-
crinology
139:4626–4633.
101. Oppenheimer JH, Schwartz HL, Surks MI 1974 Tissue dif-
ferences in the concentration of triiodothyronine nuclear
binding sites in the rat: liver, kidney, pituitary, heart, brain,
spleen, and testis. Endocrinology 95:897–903.
102. Visser WE, Friesema EC, Jansen J, Visser TJ 2007 Thyroid
hormone transport by monocarboxylate transporters. Best
Pract Res Clin Endocrinol Metab 21:223–236.
103. Braun D, Wirth EK, Schweizer, U 2010 Thyroid hormone
transporters in the brain. Rev Neurosci 21:173–186.
104. Refetoff S, Dumitrescu AM 2007 Syndromes of reduced
sensitivity to thyroid hormone: genetic defects in hormone
receptors, cell transporters and deiodination. Best Pract Res
Clin Endocrinol Metab 21: 277–305.
105. Visser WE, Friesema EC, Visser, TJ 2011 Minireview: thy-
roid hormone transporters: the knowns and the unknowns.
Mol Endocrinol 25:1–14.
106. Heuer H, Visser TJ 2009 Minireview: Pathophysiological
importance of thyroid hormone transporters. En-
docrinology 150:1078–1083.
107. Friesema EC, Grueters A, Biebermann H, Krude H, von
Moers A, Reeser M, Barrett TG, Mancilla EE, Svensson J,
Kester MH, Kuiper GG, Balkassmi S, Uitterlinden AG,
Koehrle J, Rodien P, Halestrap AP, Visser TJ 2004 Asso-
ciation between mutations in a thyroid hormone trans-
porter and severe X-linked psychomotor retardation.
Lancet 364:1435–1437.
146 BIANCO ET AL.
108. Dumitrescu AM, Liao XH, Best TB, Brockmann K, Refetoff S
2004 A novel syndrome combining thyroid and neurological
abnormalities is associated with mutations in a mono-
carboxylate transporter gene. Am J Hum Genet 74:168–175.
109. 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. Mol Endocrinol 22:1357–1369.
110. Westholm DE, Salo DR, Viken KJ, Rumbley JN, Anderson
GW 2009 The blood-brain barrier thyroxine transporter
organic anion-transporting polypeptide 1c1 displays atyp-
ical transport kinetics. Endocrinology 150:5153–5162.
111. Hagenbuch B, Gui C 2008 Xenobiotic transporters of the
human organic anion transporting polypeptides (OATP)
family. Xenobiotica 38:778–801.
112. Kinne A, Roth S, Biebermann H, Kohrle J, Gruters A,
Schweizer U 2009 Surface translocation and tri-iodothyronine
uptake of mutant MCT8 proteins are cell type-dependent.
J Mol Endocrinol 43:263–271.
113. Westholm DE, Stenehjem DD, Rumbley JN, Drewes LR,
Anderson GW 2009 Competitive inhibition of organic
anion transporting polypeptide 1c1-mediated thyroxine
transport by the fenamate class of nonsteroidal antiin-
flammatory drugs. Endocrinology 150:1025–1032.
114. Friesema EC, Kuiper GG, Jansen J, Visser TJ, Kester MH
2006 Thyroid hormone transport by the human mono-
carboxylate transporter 8 and its rate-limiting role in in-
tracellular metabolism. Mol Endocrinol 20:2761–2772.
115. Suzuki S, Mori J, Hashizume K 2007 mu-Crystallin, a
NADPH-dependent T(3)-binding protein in cytosol. Trends
Endocrinol Metab 18:286–289.
116. Visser WE, Wong WS, van Mullem AA, Friesema EC,
Geyer J, Visser, T.J. Study of the transport of thyroid hor-
mone by transporters of the SLC10 family. Mol Cell En-
docrinol 315:138–145.
117. [Deleted.]
118. Friesema EC, Ganguly S, Abdalla A, Manning Fox JE,
Halestrap AP, Visser TJ 2003 Identification of mono-
carboxylate transporter 8 as a specific thyroid hormone
transporter. J Biol Chem 278:40128–40135.
119. Sugiyama D, Kusuhara H, Shitara Y, Abe T, Sugiyama Y
2002 Effect of 17 beta-estradiol-D-17 beta-glucuronide on
the rat organic anion transporting polypeptide 2-mediated
transport differs depending on substrates. Drug Metab
Dispos 30:220–223.
120. Tracy TS 2003 Atypical enzyme kinetics: their effect on
in vitro-in vivo pharmacokinetic predictions and drug in-
teractions. Curr Drug Metab 4:341–346.
121. Hutzler JM, Tracy TS 2002 Atypical kinetic profiles in drug
metabolism reactions. Drug Metab Dispos 30:355–362.
122. Westholm DE, Marold JD, Viken KJ, Duerst AH, Anderson
GW, Rumbley JN 2010 Evidence of evolutionary conser-
vation of function between the thyroxine transporter
Oatp1c1 and major facilitator superfamily members. En-
docrinology 151:5941–5951.
123. Regina A, Morchoisne S, Borson ND, McCall AL, Drewes
LR, Roux F 2001 Factor(s) released by glucose-deprived
astrocytes enhance glucose transporter expression and ac-
tivity in rat brain endothelial cells. Biochim Biophys Acta
1540:233–242.
124. Braun D, Kinne A, Brauer AU, Sapin R, Klein MO, Kohrle J,
Wirth EK, Schweizer, U 2011 Developmental and cell type-
specific expression of thyroid hormone transporters in the
mouse brain and in primary brain cells. Glia 59:463–471.
125. Visser WE, Swagemakers SM, Ozgur Z, Schot R, Verheijen
FW, van Ijcken WF, van der Spek PJ, Visser TJ 2010 Tran-
scriptional profiling of fibroblasts from patients with mu-
tations in MCT8 and comparative analysis with the human
brain transcriptome. Hum Mol Genet 19:4189–4200.
126. Heuer H, Visser TJ 2013 The pathophysiological conse-
quences of thyroid hormone transporter deficiencies: insights
from mouse models. Biochim Biophys Acta 1830:3974–3978.
127. Morte B, Ceballos A, Diez D, Grijota-Martinez C, Dumi-
trescu AM, Di Cosmo C, Galton VA, Refetoff S, Bernal J
2010 Thyroid hormone-regulated mouse cerebral cortex
genes are differentially dependent on the source of the
hormone: a study in monocarboxylate transporter-8- and
deiodinase-2-deficient mice. Endocrinology 151:2381–2387.
128. Smith QR, Allen DD 2003 In situ brain perfusion technique.
Methods Mol Med 89:209–218.
129. Oppenheimer JH, Koerner D, Schwartz HL, Surks MI 1972
Specific nuclear triiodothyronine binding sites in rat liver
and kidney. J Clin Endocrinol Metab 35:330–333.
130. Leonard JL, Visser TJ 1986 Biochemistry of deiodination. In:
Hennemann G (ed) Thyroid Hormone Metabolism. Marcel
Dekker, New York, pp 189–229.
131. Kaplan MM 1986 Regulatory influences on iodothyronine
deiodination in animal tissues. In: Hennemann G (ed)
Thyroid Hormone Metabolism. Marcel Dekker, New York,
pp 231–253.
132. Bianco AC, Salvatore D, Gereben B, Berry MJ, Larsen PR
2002 Biochemistry, cellular and molecular biology, and
physiological roles of the iodothyronine selenodeiodinases.
Endocr Rev 23:38–89.
133. Callebaut I, Curcio-Morelli C, Mornon JP, Gereben B,
Buettner C, Huang S, Castro B, Fonseca TL, Harney JW,
Larsen PR, Bianco AC 2003 The iodothyronine seleno-
deiodinases are thioredoxin-fold family proteins containing
a glycoside hydrolase clan GH-A-like structure. J Biol
Chem 278:36887–36896.
134. Sagar GD, Gereben B, Callebaut I, Mornon JP, Zeo
¨
ld A,
Curcio-Morelli C, Harney JW, Luongo C, Mulcahey MA,
Larsen PR, Huang SA, Bianco AC 2008 The thyroid hormone-
inactivating deiodinase functions as a homodimer. Mol
Endocrinol 22:1382–1393.
135. Curcio-Morelli C, Gereben B, Zavacki AM, Kim BW,
Huang S, Harney JW, Larsen PR, Bianco AC 2003 In vivo
dimerization of types 1, 2, and 3 iodothyronine seleno-
deiodinases. Endocrinology 144:937–946.
136. Leonard JL, Larsen PR 1985 Thyroid hormone metabolism in
primary cultures of fetal rat brain cells. Brain Res 327:1–13.
137. van der Heide SM, Visser TJ, Everts ME, Klaren PH 2002
Metabolism of thyroid hormones in cultured cardiac fi-
broblasts of neonatal rats. J Endocrinol 174:111–119.
138. Albright EC, Larson FC, Tust RH 1954 In vitro conversion
of thyroxin to triiodothyronine by kidney slices. Proc Soc
Exp Biol Med 86:137–140.
139. Albright EC, Larson FC 1959 Metabolism of L-thyroxine by
human tissue slices. J Clin Invest 38:1899–1903.
140. Visser TJ, Does-Tobe I, Docter R, Hennemann G 1976
Subcellular localization of a rat liver enzyme converting
thyroxine into tri-iodothyronine and possible involvement
of essential thiol groups. Biochem J 157:479–482.
141. Harris ARC, Fang SL, Hinerfeld L, Braverman LE, Vagen-
akis AG 1979 The role of sulfhydryl groups on the impaired
hepatic 3¢,3,5-triiodothyronine generation from thyroxine
in the hypothyroid, starved, fetal and neonatal rodent.
J Clin Invest 63:516–524.
INVESTIGATING THYROID HORMONE ECONOMY AND ACTION 147
142. Verhoelst CH, Roelens SA, Darras VM 2005 Role of spa-
tiotemporal expression of iodothyronine deiodinase pro-
teins in cerebellar cell organization. Brain Res Bull 67:
196–202.
143. Kim SW, Harney JW, Larsen PR 1998 Studies of the hor-
monal regulation of type 2 5¢-iodothyronine deiodinase
messenger ribonucleic acid in pituitary tumor cells using
semiquantitative reverse transcription-polymerase chain
reaction. Endocrinology 139:4895–4905.
144. Sharifi J, St Germain DL 1992 The cDNA for the type I
iodothyronine 5¢-deiodinase encodes an enzyme manifest-
ing both high Km and low Km activity. Evidence that rat
liver and kidney contain a single enzyme which converts
thyroxine to 3,5,3¢-triiodothyronine. J Biol Chem 267:12539–
12544.
145. Berry MJ, Maia AL, Kieffer JD, Harney JW, Larsen PR 1992
Substitution of cysteine for selenocysteine in type I iodo-
thyronine deiodinase reduces the catalytic efficiency of the
protein but enhances its translation. Endocrinology
131:1848–1852.
146. Kuiper GG, Klootwijk W, Visser TJ 2002 Substitution of
cysteine for a conserved alanine residue in the catalytic
center of type II iodothyronine deiodinase alters interaction
with reducing cofactor. Endocrinology 143:1190–1198.
147. Kwakkel J, Chassande O, van Beeren HC, Fliers E, Wier-
singa WM, Boelen A 2010 Thyroid hormone receptor
{alpha} modulates lipopolysaccharide-induced changes in
peripheral thyroid hormone metabolism. Endocrinology
151:1959–1969.
148. Leonard JL, Mellen SA, Larsen PR 1983 Thyroxine 5¢-
deiodinase activity in brown adipose tissue. Endocrinology
112:1153–1155.
149. Galton VA, Martinez E, Hernandez A, St Germain EA,
Bates JM, St Germain DL 2001 The type 2 iodothyronine
deiodinase is expressed in the rat uterus and induced
during pregnancy. Endocrinology 142:2123–2128.
150. Galton VA, Hiebert A 1987 Hepatic iodothyronine 5-
deiodinase activity in Rana catesbeiana tadpoles at differ-
ent stages of the life cycle. Endocrinology 121:42–47.
151. Galton VA, Martinez E, Hernandez A, St Germain EA,
Bates JM, St Germain DL 1999 Pregnant rat uterus ex-
presses high levels of the type 3 iodothyronine deiodinase.
J Clin Invest 103:979–987.
152. Kuiper GG, Klootwijk W, Visser TJ 2003 Substitution of
cysteine for selenocysteine in the catalytic center of type III
iodothyronine deiodinase reduces catalytic efficiency and
alters substrate preference. Endocrinology 144:2505–2513.
153. Kalaany NY, Gauthier KC, Zavacki AM, Mammen PP,
Kitazume T, Peterson JA, Horton JD, Garry DJ, Bianco AC,
Mangelsdorf DJ 2005 LXRs regulate the balance between
fat storage and oxidation. Cell Metab 1:231–244.
154. Koopdonk-Kool JM, de Vijlder JJ, Veenboer GJ, Ris-Stalpers
C, Kok JH, Vulsma T, Boer K, Visser TJ 1996 Type II and
type III deiodinase activity in human placenta as a function
of gestational age. J Clin Endocrinol Metab 81:2154–2158.
155. Kester MH, Toussaint MJ, Punt CA, Matondo R, Aarnio
AM, Darras VM, Everts ME, de Bruin A, Visser TJ 2009
Large induction of type III deiodinase expression after
partial hepatectomy in the regenerating mouse and rat
liver. Endocrinology 150:540–545.
156. Richard K, Hume R, Kaptein E, Sanders JP, van Toor H, De
Herder WW, den Hollander JC, Krenning EP, Visser TJ
1998 Ontogeny of iodothyronine deiodinases in human
liver. J Clin Endocrinol Metab 83:2868–2874.
157. Kallo
´
I, Moha
´
csik P, Vida B, Zeo
¨
ld A, Bardo
´
czi Z, Zavacki
AM, Farkas E, Ka
´
da
´
r A, Hrabovszky E, Arrojo E Drigo R,
Dong L, Barna L, Palkovits M, Borsay BA, Herczeg L, Le-
chan RM, Bianco AC, Liposits Z, Fekete C, Gereben B 2012
A novel pathway regulates thyroid hormone availability in
rat and human hypothalamic neurosecretory neurons.
PLoS One 7:e37860.
158. Toyoda N, Berry MJ, Harney JW, Larsen PR 1995 Topo-
logical analysis of the integral membrane protein, type 1 io-
dothyronine deiodinase (D1). J Biol Chem 270:12310–12318.
159. Steinsapir J, Bianco AC, Buettner C, Harney J, Larsen PR
2000 Substrate-induced down-regulation of human type 2
deiodinase (hD2) is mediated through proteasomal degra-
dation and requires interaction with the enzyme’s active
center. Endocrinology 141:1127–1135.
160. [Deleted.]
161. Croteau W, Davey JC, Galton VA, St Germain DL 1996
Cloning of the mammalian type II iodothyronine deiodi-
nase. A selenoprotein differentially expressed and regu-
lated in human and rat brain and other tissues. J Clin Invest
98:405–417.
162. Gereben B, Goncalves C, Harney JW, Larsen PR, Bianco AC
2000 Selective proteolysis of human type 2 deiodinase: a
novel ubiquitin-proteasomal mediated mechanism for regu-
lation of hormone activation. Mol Endocrinol 14:1697–1708.
163. Kim BW, Zavacki AM, Curcio-Morelli C, Dentice M, Harney
JW, Larsen PR, Bianco AC 2003 Endoplasmic reticulum-
associated degradation of the human type 2 iodothyronine
deiodinase (D2) is mediated via an association between
mammalian UBC7 and the carboxyl region of D2. Mol En-
docrinol 17:2603–2612.
164. Sagar GD, Gereben B, Callebaut I, Mornon JP, Zeo
¨
ld A, da
Silva WS, Luongo C, Dentice M, Tente SM, Freitas BC,
Harney JW, Zavacki AM, Bianco AC 2007 Ubiquitination-
induced conformational change within the deiodinase
dimer is a switch regulating enzyme activity. Mol Cell Biol
27:4774–4783.
165. Botero D, Gereben B, Goncalves C, De Jesus LA, Harney JW,
Bianco AC 2002 Ubc6p and ubc7p are required for normal
and substrate-induced endoplasmic reticulum-associated
degradation of the human selenoprotein type 2 iodothyro-
nine monodeiodinase. Mol Endocrinol 16:1999–2007.
166. Steinsapir J, Harney J, Larsen PR 1998 Type 2 iodothyro-
nine deiodinase in rat pituitary tumor cells is inactivated in
proteasomes. J Clin Invest 102: 1895–1899.
167. Dentice M, Bandyopadhyay A, Gereben B, Callebaut I,
Christoffolete MA, Kim BW, Nissim S, Mornon JP, Zavacki
AM, Zeo
¨
ld A, Capelo LP, Curcio-Morelli C, Ribeiro R,
Harney JW, Tabin CJ, Bianco AC 2005 The Hedgehog-
inducible ubiquitin ligase subunit WSB-1 modulates
thyroid hormone activation and PTHrP secretion in the
developing growth plate. Nat Cell Biol 7:698–705.
168. Baqui M, Botero D, Gereben B, Curcio C, Harney JW, Sal-
vatore D, Sorimachi K, Larsen PR, Bianco AC 2003 Human
type 3 iodothyronine selenodeiodinase is located in the
plasma membrane and undergoes rapid internalization to
endosomes. J Biol Chem 278:1206–1211.
169. Huang SA, Mulcahey MA, Crescenzi A, Chung M, Kim B,
Barnes CA, Kuijt W, Tu HM, Harney JW, Larsen PR 2005
TGF-B promotes inactivation of extracellular thyroid hor-
mones via transcriptional stimulation of type 3 iodothyro-
nine deiodinase. Mol Endocrinol 19: 3126–3136.
170. Tu HM, Kim SW, Salvatore D, Bartha T, Legradi G, Larsen
PR, Lechan RM 1997 Regional distribution of type 2 thy-
148 BIANCO ET AL.
roxine deiodinase messenger ribonucleic acid in rat hypo-
thalamus and pituitary and its regulation by thyroid hor-
mone. Endocrinology 138:3359–3368.
171. Tu HM, Legradi G, Bartha T, Salvatore D, Lechan RM,
Larsen PR 1999 Regional expression of the type 3 iodo-
thyronine deiodinase messenger ribonucleic acid in the rat
central nervous system and its regulation by thyroid hor-
mone. Endocrinology 140:784–790.
172. Guadano-Ferraz A, Obregon MJ, St Germain DL, Bernal J
1997 The type 2 iodothyronine deiodinase is expressed
primarily in glial cells in the neonatal rat brain. Proc Natl
Acad Sci USA 94:10391–10396.
173. Freitas BC, Gereben B, Castillo M, Kallo I, Zeold A, Egri P,
Liposits Z, Zavacki AM, Maciel RM, Jo S, Singru P, Sanchez
E, Lechan RM, Bianco AC 2010 Paracrine signaling by glial
cell-derived triiodothyronine activates neuronal gene ex-
pression in the rodent brain and human cells. J Clin Invest
120:2206–2217.
174. Adlkofer F, Ramsden DB, Wusteman MC, Pegg DE, Hof-
fenberg R 1977 Metabolism of thyroid hormones by the
isolated perfused rabbit kidney. Horm Metab Res 9:
400–403.
175. Ferguson DC, Jennings AS 1983 Regulation of conversion
of thyroxine to triiodothyronine in perfused rat kidney. Am
J Physiol 245:E220–E229.
176. Ferguson DC, Hoenig H, Jennings AS 1985 Triiodothyronine
production by the perfused rat kidney is reduced by diabetes
mellitus but not by fasting. Endocrinology 117:64–70.
177. Jennings AS, Ferguson DC, Utiger RD 1979 Regulation of
the conversion of thyroxine to triiodothyronine in the per-
fused rat liver. J Clin Invest 64:1614–1623.
178. Jennings AS, Crutchfield FL, Dratman MB 1984 Effect of
hypothyroidism and hyperthyroidism on triiodothyro-
nine production in perfused rat liver. Endocrinology 114:
992–997.
179. Hidal JT, Kaplan MM 1985 Characteristics of thyroxine 5¢-
deiodination in cultured human placental cells: regulation
by iodothyronines. J Clin Invest 76:947–955.
180. Roti E, Braverman LE, Fang SL, Alex S, Emerson CH 1982
Ontogenesis of placental inner ring thyroxine deiodinase
and amniotic fluid 3,3¢,5¢-triiodothyronine concentration in
the rat. Endocrinology 111:959–963.
181. Roti E, Fang SL, Braverman LE, Emerson CH 1982 Rat
placenta is an active site of inner ring deiodination of
thyroxine and 3,3¢,5-triiodothyronine. Endocrinology 110:
34–37.
182. Schroder-van der Elst JP, van der Heide D, Morreale de
Escobar G, Obregon MJ 1998 Iodothyronine deiodinase
activities in fetal rat tissues at several levels of iodine de-
ficiency: a role for the skin in 3,5,3¢-triiodothyronine econ-
omy? Endocrinology 139:2229–2234.
183. Castro MI, Braverman LE, Alex S, Wu CF, Emerson CH
1985 Inner-ring deiodination of 3,5,3¢-triiodothyronine in
the in situ perfused guinea pig placenta. J Clin Invest
76:1921–1926.
184. Cooper E, Gibbens M, Thomas CR, Lowy C, Burke CW
1983 Conversion of thyroxine to 3,3¢,5¢-triiodothyronine in
the guinea pig placenta: in vivo studies. Endocrinology
112:1808–1815.
185. Kaplan MM, Shaw EA 1984 Type II iodothyronine 5¢
-
deiodination by human and rat placenta in vitro. J Clin
Endocrinol Metab 59:253–257.
186. Schneider MJ, Fiering SN, Thai B, Wu SY, St Germain E,
Parlow AF, St Germain DL, Galton VA 2006 Targeted
disruption of the type 1 selenodeiodinase gene (dio1) re-
sults in marked changes in thyroid hormone economy in
mice. Endocrinology 147:580–589.
187. Galton VA, Wood ET, St Germain EA, Withrow CA, Al-
drich G, St Germain GM, Clark AS, St Germain DL 2007
Thyroid hormone homeostasis and action in the type 2
deiodinase-deficient rodent brain during development.
Endocrinology 148:3080–3088.
188. Visser TJ 1996 Pathways of thyroid hormone metabolism.
Acta Medica Austriaca 23:10–16.
189. Curran PG, DeGroot LJ 1991 The effect of hepatic enzyme-
inducing drugs on thyroid hormones and the thyroid
gland. Endocr Rev 12: 135–150.
190. Peeters RP, Kester MH, Wouters PJ, Kaptein E, van Toor H,
Visser TJ, Van den Berghe G 2005 Increased thyroxine
sulfate levels in critically ill patients as a result of a de-
creased hepatic type I deiodinase activity. J Clin Endocrinol
Metab 90:6460–6465.
191. Mackenzie PI, Bock KW, Burchell B, Guillemette C,
Ikushiro S, Iyanagi T, Miners JO, Owens IS, Nebert DW
2005 Nomenclature update for the mammalian UDP gly-
cosyltransferase (UGT) gene superfamily. Pharmacogenet
Genomics 15:677–685.
192. Wu SY, Green WL, Huang WS, Hays MT, Chopra IJ 2005
Alternate pathways of thyroid hormone metabolism. Thy-
roid 15:943–958.
193. Stanley EL, Hume R, Coughtrie MW 2005 Expression
profiling of human fetal cytosolic sulfotransferases in-
volved in steroid and thyroid hormone metabolism and in
detoxification. Mol Cell Endocrinol 240:32–42.
194. Kester MH, Kaptein E, Roest TJ, van Dijk CH, Tibboel D,
Meinl W, Glatt H, Coughtrie MW, Visser TJ 1999 Char-
acterization of human iodothyronine sulfotransferases. J
Clin Endocrinol Metab 84:1357–1364.
195. Pietsch CA, Scanlan TS, Anderson RJ 2007 Thyronamines
are substrates for human liver sulfotransferases. En-
docrinology 148:1921–1927.
196. Kaptein E, van Haasteren GA, Linkels E, de Greef WJ, Visser
TJ 1997 Characterization of iodothyronine sulfotransferase
activity in rat liver. Endocrinology 138:5136–5143.
197. Kato Y, Ikushiro S, Emi Y, Tamaki S, Suzuki H, Sakaki T,
Yamada S, Degawa M 2008 Hepatic UDP-glucuronosyl-
transferases responsible for glucuronidation of thyroxine in
humans. Drug Metab Dispos 36:51–55.
198. Moreno M, Kaptein E, Goglia F, Visser TJ 1994 Rapid
glucuronidation of tri- and tetraiodothyroacetic acid to
ester glucuronides in human liver and to ether glucuro-
nides in rat liver. Endocrinology 135:1004–1009.
199. Visser TJ, Kaptein E, van Toor H, van Raaij JA, van den
Berg KJ, Joe CT, van Engelen JG, Brouwer A 1993 Glucur-
onidation of thyroid hormone in rat liver: effects of in vivo
treatment with microsomal enzyme inducers and in vitro
assay conditions. Endocrinology 133:2177–2186.
200. Frumess RD, Larsen PR 1975 Correlation of serum triio-
dothyronine (T3) and thyroxine (T4) with biologic effects of
thyroid hormone replacement in propylthiouracil- treated
rats. Metabolism 24:547–554.
201. Hervas F, Morreale de Escobar G, Escobar Del Rey F 1975
Rapid effects of single small doses of L-thyroxine and
triiodo-L-thyronine on growth hormone, as studied in the
rat by radioimmunoassy. Endocrinology 97: 91–101.
202. Larsen PR, Frumess RD 1977 Comparison of the biological
effects of thyroxine and triiodothyronine in the rat. En-
docrinology 100:980–988.
INVESTIGATING THYROID HORMONE ECONOMY AND ACTION 149
203. Clark OH, Lambert WR, Cavalieri RR, Rapoport B, Ham-
mond ME, Ingbar SH 1976 Compensatory thyroid hyper-
trophy after hemithyroidectomy in rats. Endocrinology
99:988–995.
204. Saha SK, Ohinata H, Ohno T, Kuroshima A 1998 Thermo-
genesis and fatty acid composition of brown adipose tissue in
rats rendered hyperthyroid andhypothyroidwithspecial
reference to docosahexaenoic acid. Jpn J Physiol 48:355–364.
205. Branvold DJ, Allred DR, Beckstead DJ, Kim HJ, Fillmore N,
Condon BM, Brown JD, Sudweeks SN, Thomson DM,
Winder WW 2008 Thyroid hormone effects on LKB1,
MO
2
5, phospho-AMPK, phospho-CREB, and PGC-1alpha
in rat muscle. J Appl Physiol 105:1218–1227.
206. Boughter JD Jr, Raghow S, Nelson TM, Munger SD 2005
Inbred mouse strains C57BL/6J and DBA/2J vary in sen-
sitivity to a subset of bitter stimuli. BMC Genet 6:36.
207. Chatoui H, El Hiba O, Elgot A, Gamrani H 2012 The rat
SCO responsiveness to prolonged water deprivation: im-
plication of Reissner’s fiber and serotonin system. C R Biol
335:253–260.
208. Cawthorne C, Swindell R, Stratford IJ, Dive C, Welman A
2007 Comparison of doxycycline delivery methods for Tet-
inducible gene expression in a subcutaneous xenograft
model. J Biomol Tech 18:120–123.
209. Rondeel JMM, De Greef WJ, Klootwijk W, Visser TJ 1992
Effects of hypothyroidism on hypothalamic release of
thyrotropin-releasing hormone in rats. Endocrinology 130:
651–656.
210. Solis SJ, Villalobos P, Orozco A, Delgado G, Quintanar-
Stephano A, Garcia-Solis P, Hernandez-Montiel HL, Robles-
Osorio L, Valverde RC 2010 Inhibition of intrathyroidal
dehalogenation by iodide. J Endocrinol 208:89–96.
211. Ueta CB, Olivares EL, Bianco AC 2011 Responsiveness to
thyroid hormone and to ambient temperature underlies
differences between brown adipose tissue and skeletal
muscle thermogenesis in a mouse model of diet-induced
obesity. Endocrinology 152:3571–3581.
212. Perello M, Friedman T, Paez-Espinosa V, Shen X, Stuart
RC, Nillni EA 2006 Thyroid hormones selectively regulate
the posttranslational processing of prothyrotropin-releasing
hormone in the paraventricular nucleus of the hypothala-
mus. Endocrinology 147:2705–2716.
213. Goldberg RC, Chaikoff IL, Lindsay S, Feller DD 1950 His-
topathological changes induced in the normal thyroid and
other tissues of the rat by internal radiation with various
doses of radioactive iodine. Endocrinology 46:72–90.
214. Feller DD, Chaikoff IL, Taurog A, Jones HB 1949 The
changes induced in iodine metabolism of the rat by internal
radiation of its thyroid with I131. Endocrinology 45:
464–479.
215. Goldberg RC, Chaikoff IL 1949 A simplified procedure for
thyroidectomy of the newborn rat without concomitant
parathyroidectomy. Endocrinology 45:64–70.
216. Kasuga Y, Matsubayashi S, Sakatsume Y, Akasu F, Jamie-
son C, Volpe R 1991 The effect of xenotransplantation of
human thyroid tissue following radioactive iodine-induced
thyroid ablation on thyroid function in the nude mouse.
Clin Invest Med 14:277–281.
217. Seo H, Wunderlich C, Vassart G, Refetoff S 1981 Growth
hormone responses to thyroid hormone in the neonatal rat:
resistance and anamnestic response. J Clin Invest 67:
569–574.
218. Smith PE 1930 Hypophysectomy and a replacement ther-
apy in the rat. Am J Anatomy 45:205–273.
219. Renaud S, Picard G 1964 An improved table for hypoph-
ysectomy in rat. Can J Physiol Pharmacol 42:870–872.
220. Falconi G, Rossi GL 1964 Method for placing a pituitary
graft into the evacuated pituitary capsule of the hypophy-
sectomized rat or mouse. Endocrinology 75:964–967.
221. Sato M, Yoneda S 1966 An efficient method for transauri-
cular hypophysectomy in rats. Acta Endocrinol (Copenh)
51:43–48.
222. Leung GS, Kawai M, Tai JK, Chen J, Bandiera SM, Chang
TK 2009 Developmental expression and endocrine regula-
tion of CYP1B1 in rat testis. Drug Metab Dispos 37:523–528.
223. Kimura S, Hara Y, Pineau T, Fernandez-Salguero P, Fox
CH, Ward JM, Gonzalez FJ 1996 The T/ebp null mouse:
thyroid-specific enhancer-binding protein is essential for
the organogenesis of the thyroid, lung, ventral forebrain,
and pituitary. Genes Dev 10:60–69.
224. Mansouri A, Chowdhury K, Gruss P 1998 Follicular cells of
the thyroid gland require Pax8 gene function. Nat Genet
19:87–90.
225. Adams PM, Stein SA, Palnitkar M, Anthony A, Gerrity L,
Shanklin DR 1989 Evaluation and characterization of the
hypothyroid hyt/hyt mouse. I: Somatic and behavioral
studies. Neuroendocrinology 49:138–143.
226. Stein SA, Shanklin DR, Krulich L, Roth MG, Chubb CM,
Adams PM 1989 Evaluation and characterization of the
hyt/hyt hypothyroid mouse. II. Abnormalities of TSH and
the thyroid gland. Neuroendocrinology 49:509–519.
227. Marians RC, Ng L, Blair HC, Unger P, Graves PN, Davies
TF 2002 Defining thyrotropin-dependent and -independent
steps of thyroid hormone synthesis by using thyrotropin
receptor-null mice. Proc Natl Acad Sci USA 99:15776–15781.
228. Johnson KR, Marden CC, Ward-Bailey P, Gagnon LH,
Bronson RT, Donahue LR 2007 Congenital hypothyroid-
ism, dwarfism, and hearing impairment caused by a mis-
sense mutation in the mouse dual oxidase 2 gene, Duox2.
Mol Endocrinol 21:1593–1602.
229. [Deleted.]
230. Beamer WG, Maltais LJ, DeBaets MH, Eicher EM 1987 In-
herited congenital goiter in mice. Endocrinology 120:
838–840.
231. Taylor BA, Rowe L 1987 The congenital goiter mutation is
linked to the thyroglobulin gene in the mouse. Proc Natl
Acad Sci USA 84: 1986–1990.
232. Castillo M, Hall JA, Correa-Medina M, Ueta C, Kang HW,
Cohen DE, Bianco AC 2011 Disruption of thyroid hormone
activation in type 2 deiodinase knockout mice causes obe-
sity with glucose intolerance and liver steatosis only at
thermoneutrality. Diabetes 60:1082–1089.
233. Watanabe M, Houten SM, Mataki C, Christoffolete MA,
Kim BW, Sato H, Messaddeq N, Harney JW, Ezaki O,
Kodama T, Schoonjans K, Bianco AC, Auwerx J 2006 Bile
acids induce energy expenditure by promoting intracellular
thyroid hormone activation. Nature 439:484–489.
234. de Jesus LA, Carvalho SD, Ribeiro MO, Schneider M, Kim
SW, Harney JW, Larsen PR, Bianco AC 2001 The type 2
iodothyronine deiodinase is essential for adaptive thermo-
genesis in brown adipose tissue. J Clin Invest 108:1379–1385.
235. Rosene ML, Wittmann G, Arrojo e Drigo R, Singru PS,
Lechan RM, Bianco AC 2010 Inhibition of the type 2 io-
dothyronine deiodinase underlies the elevated plasma TSH
associated with amiodarone treatment. Endocrinology
151:5961–5970.
236. Reiter RJ, Klaus S, Ebbinghaus C, Heldmaier G, Redlin U,
Ricquier D, Vaughan MK, Steinlechner S 1990 Inhibition of
150 BIANCO ET AL.
5¢-deiodination of thyroxine suppresses the cold-induced
increase in brown adipose tissue messenger ribonucleic
acid for mitochondrial uncoupling protein without influ-
encing lipoprotein lipase activity. Endocrinology 126:2550–
2554.
237. Bianco AC, Carvalho SD, Carvalho CR, Rabelo R, Moriscot
AS 1998 Thyroxine 5¢-deiodination mediates norepinephrine-
induced lipogenesis in dispersed brown adipocytes. En-
docrinology 139:571–578.
238. Pol CJ, Muller A, Zuidwijk MJ, van Deel ED, Kaptein E,
Saba A, Marchini M, Zucchi R, Visser TJ, Paulus WJ,
Duncker DJ, Simonides WS 2011 Left-ventricular re-
modeling after myocardial infarction is associated with a
cardiomyocyte-specific hypothyroid condition. Endocrin-
ology 152:669–679.
239. Medina MC, Molina J, Gadea Y, Fachado A, Murillo M,
Simovic G, Pileggi A, Herna
´
ndez A, Edlund H, Bianco AC
2011 The thyroid hormone-inactivating type III deiodinase
is expressed in mouse and human beta-cells and its tar-
geted inactivation impairs insulin secretion. Endocrinology
152:3717–3727.
240. [Deleted.]
241. Olivares EL, Marassi MP, Fortunato RS, da Silva AC,
Costa-e-Sousa RH, Arau
´
jo IG, Mattos EC, Masuda MO,
Mulcahey MA, Huang SA, Bianco AC, Carvalho DP 2007
Thyroid function disturbance and type 3 iodothyronine
deiodinase induction after myocardial infarction in rats a
time course study. Endocrinology 148: 4786–4792.
242. Abrams GM, Larsen PR 1973 Triiodothyronine and thy-
roxine in the serum and thyroid glands of iodine-deficient
rats. J Clin Invest 52:2522–2531.
243. Escobar-Morreale HF, Obregon MJ, Escobar del Rey F,
Morreale de Escobar G 1995 Replacement therapy for hy-
pothyroidism with thyroxine alone does not ensure eu-
thyroidism in all tissues, as studied in thyroidectomized
rats. J Clin Invest 96:2828–2838.
244. Escobar-Morreale HF, Escobar del Rey F, Obregon MJ,
Morreale de Escobar G 1996 Only the combined treatment
with thyroxine and triiodothyronine ensures euthyroidism
in all tissues of the thyroidectomized rat. Endocrinology
137:2490–2502.
245. Frumess RD, Larsen PR 1975 The effect of inhibiting triio-
dothyronine (T3) production from thyroxine (T4) by pro-
pylthiouracil (PTU) on the physiological activity of T4 in
thyroidectomized rats. In: Harland WA, Orr JS (eds) Thy-
roid Hormone Metabolism. Academic Press, Waltham,
MA, pp 125–137.
246. Larsen PR, Frumess RD 1976 Comparison of the effects of
triiodothyronine (T3) and thyroxine (T4) on serum TSH and
hepatic mitochondrial glycerophosphate dehydrogenase
(alpha-GPD) activities in thyroidectomized hypothyroid
rats. In: Robbins J, Braverman LE (eds) Thyroid Research:
Proceedings of the Seventh International Thyroid Con-
ference, Boston, Massachusetts, June 9–13, 1975. Excerpta
Medica, Amsterdam, pp 21–24.
247. Deol MS 1973 An experimental approach to the understand-
ing and treatment of hereditary syndromes with congenital
deafness and hypothyroidism. J Med Genet 10:235–242.
248. He
´
bert R, Langlois JM, Dussault JH 1985 Permanent defects
in rat peripheral auditory function following perinatal hy-
pothyroidism: determination of a critical period. Brain Res
23:161–170.
249. Sprenkle PM, McGee J, Bertoni JM, Walsh EJ 2001 Pre-
vention of auditory dysfunction in hypothyroid Tshr mu-
tant mice by thyroxin treatment during development. J
Assoc Res Otolaryngol 2:348–361.
250. Sprenkle PM, McGee J, Bertoni JM, Walsh EJ 2001 Devel-
opment of auditory brainstem responses (ABRs) in Tshr
mutant mice derived from euthyroid and hypothyroid
dams. J Assoc Res Otolaryngol 2:330–347.
251. Sprenkle PM, McGee J, Bertoni JM, Walsh EJ 2001 Con-
sequences of hypothyroidism on auditory system function
in Tshr mutant (hyt) mice. J Assoc Res Otolaryngol 2:312–
329.
252. Antonica F, Kasprzyk DF, Opitz R, Iacovino M, Liao XH,
Dumitrescu AM, Refetoff S, Peremans K, Manto M, Kyba
M, Costagliola S 2012 Generation of functional thyroid
from embryonic stem cells. Nature 491:66–71.
253. Ma R, Latif R, Davies TF 2013 Thyroid follicle formation
and thyroglobulin expression in multipotent endodermal
stem cells. Thyroid 23:385–391.
254. Moeller LC, Wardrip C, Niekrasz M, Refetoff S, Weiss RE
2009 Comparison of thyroidectomized calf serum and
stripped serum for the study of thyroid hormone action in
human skin fibroblasts in vitro. Thyroid 19:639–644.
255. Samuels HH, Stanley F, Casanova J 1979 Depletion of L-
3,5,3¢-triiodothyronine and L-thyroxine in euthyroid calf
serum for use in cell culture studies of the action of thyroid
hormone. Endocrinology 105:80–85.
256. Hollenberg AN, Monden T, Madura JP, Lee K, Wondisford
FE 1996 Function of nuclear co-repressor protein on thyroid
hormone response elements is regulated by the receptor A/
B domain. J Biol Chem 271:28516–28520.
257. Cao Z, West C, Norton-Wenzel CS, Rej R, Davis FB, Davis
PJ 2009 Effects of resin or charcoal treatment on fetal bovine
serum and bovine calf serum. Endocr Res 34:101–108.
258. Muller A, Zuidwijk MJ, van Hardeveld C 1993 Effects of
thyroid hormone on growth and differentiation of L6
muscle cells. BAM 3:59–68.
259. Christoffolete MA, Ribeiro R, Singru P, Fekete C, da Silva WS,
Gordon DF, Huang SA, Crescenzi A, Harney JW, Ridgway
EC, Larsen PR, Lechan RM, Bianco AC 2006 Atypical ex-
pression of type 2 iodothyronine deiodinase in thyrotrophs
explains the thyroxine-mediated pituitary thyrotropin feed-
back mechanism. Endocrinology 147:1735–1743.
260. Shah V, Nguyen P, Nguyen NH, Togashi M, Scanlan TS,
Baxter JD, Webb P 2008 Complex actions of thyroid hor-
mone receptor antagonist NH-3 on gene promoters in dif-
ferent cell lines. Mol Cell Endocrinol 296:69–77.
261. Grover GJ, Dunn C, Nguyen NH, Boulet J, Dong G, Do-
mogauer J, Barbounis P, Scanlan TS 2007 Pharmacological
profile of the thyroid hormone receptor antagonist NH3 in
rats. J Pharmacol Exp Ther 322:385–390.
262. Nguyen NH, Apriletti JW, Cunha Lima ST, Webb P, Baxter
JD, Scanlan TS 2002 Rational design and synthesis of a
novel thyroid hormone antagonist that blocks coactivator
recruitment. J Med Chem 45:3310–3320.
263. Lim W, Nguyen NH, Yang HY, Scanlan TS, Furlow JD 2002
A thyroid hormone antagonist that inhibits thyroid hor-
mone action in vivo. J Biol Chem 277:35664–35670.
264. Baxter JD, Goede P, Apriletti JW, West BL, Feng W, Mell-
strom K, Fletterick RJ, Wagner RL, Kushner PJ, Ribeiro RC,
Webb P, Scanlan TS, Nilsson S 2002 Structure-based design
and synthesis of a thyroid hormone receptor (TR) antago-
nist. Endocrinology 143:517–524.
265. Yoshihara HA, Apriletti JW, Baxter JD, Scanlan TS 2001 A
designed antagonist of the thyroid hormone receptor.
Bioorg Med Chem Lett 11:2821–2825.
INVESTIGATING THYROID HORMONE ECONOMY AND ACTION 151
266. Seelig S, Jump DB, Towle HC, Liaw C, Mariash CN,
Schwartz HL, Oppenheimer JH 1982 Paradoxical effects of
cycloheximide on the ultra-rapid induction of two hepatic
mRNA sequences by triiodothyronine (T3). Endocrinology
110:671–673.
267. Bianco AC, Silva JE 1987 Intracellular conversion of thy-
roxine to triiodothyronine is required for the optimal
thermogenic function of brown adipose tissue. J Clin Invest
79:295–300.
268. Bianco AC, Silva JE 1987 Optimal response of key enzymes
and uncoupling protein to cold in BAT depends on local T
3
generation. Am J Physiol 253:E255–E263.
269. Bianco AC, Sheng XY, Silva JE 1988 Triiodothyronine am-
plifies norepinephrine stimulation of uncoupling protein
gene transcription by a mechanism not requiring protein
synthesis. J Biol Chem 263:18168–18175.
270. [Deleted.]
271. Carvalho SD, Kimura ET, Bianco AC, Silva JE 1991 Central
role of brown adipose tissue thyroxine 5¢-deiodinase on
thyroid hormone-dependent thermogenic response to cold.
Endocrinology 128:2149–2159.
272. [Deleted.]
273. Carvalho-Bianco SD, Kim BW, Zhang JX, Harney JW, Ri-
beiro RS, Gereben B, Bianco AC, Mende U, Larsen PR 2004
Chronic cardiac-specific thyrotoxicosis increases myocar-
dial beta-adrenergic responsiveness. Mol Endocrinol
18:1840–1849.
274. Trivieri MG, Oudit GY, Sah R, Kerfant BG, Sun H, Gramolini
AO, Pan Y, Wickenden AD, Croteau W, Morreale de Escobar
G, Pekhletski R, St Germain D, Maclennan DH, Backx PH
2006 Cardiac-specific elevations in thyroid hormone enhance
contractility and prevent pressure overload-induced cardiac
dysfunction. Proc Natl Acad Sci USA 103:6043–6048.
275. Hernandez A, Martinez ME, Fiering S, Galton VA, St
Germain D 2006 Type 3 deiodinase is critical for the mat-
uration and function of the thyroid axis. J Clin Invest
116:476–484.
276. Hernandez A, Martinez ME, Liao XH, Van Sande J, Refetoff
S, Galton VA, St Germain DL 2007 Type 3 deiodinase de-
ficiency results in functional abnormalities at multiple
levels of the thyroid axis. Endocrinology 148:5680–5687.
277. Ng L, Hernandez A, He W, Ren T, Srinivas M, Ma M,
Galton VA, St Germain DL, Forrest D 2009 A protective
role for type 3 deiodinase, a thyroid hormone-inactivating
enzyme, in cochlear development and auditory function.
Endocrinology 150:1952–1960.
278. Hernandez A, Quignodon L, Martinez ME, Flamant F, St
Germain DL 2010 Type 3 deiodinase deficiency causes
spatial and temporal alterations in brain T
3
signaling that
are dissociated from serum thyroid hormone levels. En-
docrinology 151:5550–5558.
279. Ng L, Lyubarsky A, Nikonov SS, Ma M, Srinivas M, Kefas
B, St Germain DL, Hernandez A, Pugh EN Jr, Forrest D
2010 Type 3 deiodinase, a thyroid-hormone-inactivating
enzyme, controls survival and maturation of cone photo-
receptors. J Neurosci 30:3347–3357.
280. [Deleted.]
281. Murata Y, Ceccarelli P, Refetoff S, Horwitz AL, Matsui N 1987
Thyroid hormone inhibits fibronectin synthesis by cultured
humanskinfibroblasts.JClinEndocrinolMetab64:334–339.
282. Miura M, Tanaka K, Komatsu Y, Suda M, Yasoda A, Sa-
kuma Y, Ozasa A, Nakao K 2002 Thyroid hormones pro-
mote chondrocyte differentiation in mouse ATDC5 cells
and stimulate endochondral ossification in fetal mouse
tibias through iodothyronine deiodinases in the growth
plate. J Bone Miner Res 17:443–454.
283. Al-Jubouri MA, Inkster GD, Nee PA, Andrews FJ 2006
Thyrotoxicosis presenting as hypokalaemic paralysis and
hyperlactataemia in an oriental man. Ann Clin Biochem
43:323–325.
284. Ho J, Jackson R, Johnson D 2011 Massive levothyroxine
ingestion in a pediatric patient: case report and discussion.
CJEM 13:165–168.
285. Brent GA 2000 Tissue-specific actions of thyroid hormone:
insights from animal models. Rev Endocr Metab Disord
1:27–33.
286. Forrest D, Vennstrom B 2000 Functions of thyroid hormone
receptors in mice. Thyroid 10:41–52.
287. Pennock GD, Raya TE, Bahl JJ, Goldman S, Morkin E 1992
Cardiac effects of 3,5-diiodothyropropionic acid, a thyroid
hormone analog with inotropic selectivity. J Pharmacol Exp
Ther 263:163–169.
288. Mahaffey KW, Raya TE, Pennock GD, Morkin E, Goldman
S 1995 Left ventricular performance and remodeling in
rabbits after myocardial infarction. Effects of a thyroid
hormone analogue. Circulation 91:794–801.
289. Pennock GD, Raya TE, Bahl JJ, Goldman S, Morkin E 1993
Combination treatment with captopril and the thyroid
hormone analogue 3,5-diiodothyropropionic acid. A new
approach to improving left ventricular performance in
heart failure. Circulation 88:1289–1298.
290. Abohashem-Aly AA, Meng X, Li J, Sadaria MR, Ao L,
Wennergren J, Fullerton DA, Raeburn CD 2011 DITPA, a
thyroid hormone analog, reduces infarct size and attenu-
ates the inflammatory response following myocardial is-
chemia. J Surg Res 171:379–385.
291. Tomanek RJ, Zimmerman MB, Suvarna PR, Morkin E,
Pennock GD, Goldman S 1998 A thyroid hormone analog
stimulates angiogenesis in the post-infarcted rat heart. J
Mol Cell Cardiol 30:923–932.
292. Ladenson PW, McCarren M, Morkin E, Edson RG, Shih
MC, Warren SR, Barnhill JG, Churby L, Thai H, O’Brien T,
Anand I, Warner A, Hattler B, Dunlap M, Erikson J,
Goldman S 2010 Effects of the thyromimetic agent diio-
dothyropropionic acid on body weight, body mass index,
and serum lipoproteins: a pilot prospective, randomized,
controlled study. J Clin Endocrinol Metab 95:1349–1354.
293. Goldman S, McCarren M, Morkin E, Ladenson PW, Edson
R, Warren S, Ohm J, Thai H, Churby L, Barnhill J, O’Brien
T, Anand I, Warner A, Hattler B, Dunlap M, Erikson J, Shih
MC, Lavori P 2009 DITPA (3,5-diiodothyropropionic
acid), a thyroid hormone analog to treat heart failure:
phase II trial veterans affairs cooperative study. Circulation
119:3093–3100.
294. Hadi NR, Al-amran FG, Hussein AA 2011 Effects of thyroid
hormone analogue and a leukotrienes pathway-blocker on
renal ischemia/reperfusion injury in mice. BMC Nephrol
12:70.
295. Talukder MA, Yang F, Nishijima Y, Chen CA, Xie L, Ma-
hamud SD, Kalyanasundaram A, Bonagura JD, Periasamy
M, Zweier JL 2011 Detrimental effects of thyroid hormone
analog DITPA in the mouse heart: increased mortality with
in vivo acute myocardial ischemia-reperfusion. Am J Phy-
siol Heart Circ Physiol 300:H702–H711.
296. Wagner RL, Huber BR, Shiau AK, Kelly A, Cunha Lima ST,
Scanlan TS, Apriletti JW, Baxter JD, West BL, Fletterick RJ
2001 Hormone selectivity in thyroid hormone receptors.
Mol Endocrinol 15:398–410.
152 BIANCO ET AL.
297. Schueler PA, Schwartz HL, Strait KA, Mariash CN, Op-
penheimer JH 1990 Binding of 3,5,3¢-triiodothyronine (T
3
)
and its analogs to the in vitro translational products of
c-erbA protooncogenes: differences in the affinity of the a-
and b-forms for the acetic acid analog and failure of the
human testis and kidney a-2 products to bind T
3
. Mol
Endocrinol 4:227–234.
298. Martinez L, Nascimento AS, Nunes FM, Phillips K, Apar-
icio R, Dias SM, Figueira AC, Lin JH, Nguyen P, Apriletti
JW, Neves FA, Baxter JD, Webb P, Skaf MS, Polikarpov I
2009 Gaining ligand selectivity in thyroid hormone recep-
tors via entropy. Proc Natl Acad Sci USA 106:20717–20722.
299. Massol J, Martin P, Soubrie P, Simon P 1987 Triiodothyro-
acetic acid-induced reversal of learned helplessness in rats.
Eur J Pharmacol 134:345–348.
300. Medina-Gomez G, Hernandez A, Calvo RM, Martin E,
Obregon MJ 2003 Potent thermogenic action of triio-
dothyroacetic acid in brown adipocytes. Cell Mol Life Sci
60:1957–1967.
301. Medina-Gomez G, Calvo RM, Obregon MJ 2008 Thermo-
genic effect of triiodothyroacetic acid at low doses in rat
adipose tissue without adverse side effects in the thyroid
axis. Am J Physiol Endocrinol Metab 294:E688–697.
302. Liang H, Juge-Aubry CE, O’Connell M, Burger AG 1997
Organ-specific effects of 3,5,3¢-triiodothyroacetic acid in
rats. Eur J Endocrinol 137:537–544.
303. Salmela PI, Wide L, Juustila H, Ruokonen A 1988 Effects of
thyroid hormones (T4, T3), bromocriptine and Triac on
inappropriate TSH hypersecretion. Clin Endocrinol (Oxf)
28:497–507.
304. Beck-Peccoz P, Sartorio A, De Medici C, Grugni G, Mor-
abito F, Faglia G 1988 Dissociated thyromimetic effects of 3,
5,3¢-triiodothyroacetic acid (TRIAC) at the pituitary and
peripheral tissue levels. J Endocrinol Invest 11:113–118.
305. Mueller-Gaertner HW, Schneider C 1988 3,5,3¢-Triiodo-
thyroacetic acid minimizes the pituitary thyrotrophin se-
cretion in patients on levo-thyroxine therapy after ablative
therapy for differentiated thyroid carcinoma. Clin Endo-
crinol (Oxf) 28:345–351.
306. Sherman SI, Ladenson PW 1992 Organ-specific effects of tir-
atricol: a thyroid hormone analog with hepatic, not pituitary,
superagonist effects. J Clin Endocrinol Metab 75:901–905.
307. Sherman SI, Ringel MD, Smith MJ, Kopelen HA, Zoghbi
WA, Ladenson PW 1997 Augmented hepatic and skeletal
thyromimetic effects of tiratricol in comparison with le-
vothyroxine. J Clin Endocrinol Metab 82:2153–2158.
308. Chiellini G, Apriletti JW, al Yoshihara H, Baxter JD, Ribeiro
RC, Scanlan TS 1998 A high-affinity subtype-selective ag-
onist ligand for the thyroid hormone receptor. Chem Biol
5:299–306.
309. Ribeiro RC, Apriletti JW, Wagner RL, Feng W, Kushner PJ,
Nilsson S, Scanlan TS, West BL, Fletterick RJ, Baxter JD
1998 X-ray crystallographic and functional studies of thy-
roid hormone receptor. J Steroid Biochem Mol Biol 65:133–
141.
310. Trost SU, Swanson E, Gloss B, Wang-Iverson DB, Zhang H,
Volodarsky T, Grover GJ, Baxter JD, Chiellini G, Scanlan
TS, Dillmann WH 2000 The thyroid hormone receptor-beta-
selective agonist GC-1 differentially affects plasma lipids
and cardiac activity. Endocrinology 141:3057–3064.
311. Ribeiro MO, Carvalho SD, Schultz JJ, Chiellini G, Scanlan TS,
Bianco AC, Brent GA 2001 Thyroid hormone–sympathetic
interaction and adaptive thermogenesis are thyroid hor-
mone receptor isoform–specific. J Clin Invest 108:97–105.
312. Freitas FRS, Zorn T, Labatte C, Scanlan TS, Brent GA,
Moriscot AS, Bianco AC, Gouveia CHA 2002 Effects of the
thyroid hormone receptor beta (TRb)-selective compound
GC-1 on bone development of Wistar rats. In: 74th Annual
Meeting of the American Thyroid Association. American
Thyroid Association, Los Angeles, CA.
313. Freitas FR, Moriscot AS, Jorgetti V, Soares AG, Passarelli
M, Scanlan TS, Brent GA, Bianco AC, Gouveia CH 2003
Spared bone mass in rats treated with thyroid hormone
receptor TR beta-selective compound GC-1. Am J Physiol
Endocrinol Metab 285:E1135–E1141.
314. Manzano J, Morte B, Scanlan TS, Bernal J 2003 Differential
effects of triiodothyronine and the thyroid hormone re-
ceptor beta-specific agonist GC-1 on thyroid hormone tar-
get genes in the brain. Endocrinology 144:5480–5487.
315. Grover GJ, Egan DM, Sleph PG, Beehler BC, Chiellini G,
Nguyen NH, Baxter JD, Scanlan TS 2004 Effects of the
thyroid hormone receptor agonist GC-1 on metabolic rate
and cholesterol in rats and primates: selective actions
relative to 3,5,3¢-triiodo-L-thyronine. Endocrinology 145:
1656–1661.
316. Martinez de Mena R, Scanlan TS, Obregon, MJ 2010 The T
3
receptor beta1 isoform regulates UCP1 and D2 deiodinase
in rat brown adipocytes. Endocrinology 151:5074–5083.
317. Bryzgalova G, Effendic S, Khan A, Rehnmark S, Barbounis
P, Boulet J, Dong G, Singh R, Shapses S, Malm J, Webb P,
Baxter JD, Grover GJ 2008 Anti-obesity, anti-diabetic, and
lipid lowering effects of the thyroid receptor beta subtype
selective agonist KB-141. J Steroid Biochem Mol Biol
111:262–267.
318. Grover GJ, Mellstro
¨
m K, Malm J 2007 Therapeutic potential
for thyroid hormone receptor-beta selective agonists for
treating obesity, hyperlipidemia and diabetes. Curr Vasc
Pharmacol 5:141–154.
319. Erion MD, Cable EE, Ito BR, Jiang H, Fujitaki JM, Finn PD,
Zhang BH, Hou J, Boyer SH, van Poelje PD, Linemeyer DL
2007 Targeting thyroid hormone receptor-beta agonists to
the liver reduces cholesterol and triglycerides and im-
proves the therapeutic index. Proc Natl Acad Sci USA
104:15490–15495.
320. Grover GJ, Mellstrom K, Ye L, Malm J, Li YL, Bladh LG,
Sleph PG, Smith MA, George R, Vennstro
¨
m B, Mookhtiar
K, Horvath R, Speelman J, Egan D, Baxter JD 2003 Selective
thyroid hormone receptor-beta activation: a strategy for
reduction of weight, cholesterol, and lipoprotein (a) with
reduced cardiovascular liability. Proc Natl Acad Sci USA
100:10067–10072.
321. Berkenstam A, Kristensen J, Mellstro
¨
m K, Carlsson B,
Malm J, Rehnmark S, Garg N, Andersson CM, Rudling M,
Sjo
¨
berg F, Angelin B, Baxter JD 2008 The thyroid hormone
mimetic compound KB2115 lowers plasma LDL cholesterol
and stimulates bile acid synthesis without cardiac effects in
humans. Proc Natl Acad Sci USA 105:663–667.
322. Ladenson PW, Kristensen JD, Ridgway EC, Olsson AG,
Carlsson B, Klein I, Baxter JD, Angelin B 2010 Use of the
thyroid hormone analogue eprotirome in statin-treated
dyslipidemia. N Engl J Med 362:906–916.
323. Karo Bio AB 2012 Karo Bio terminates the eprotirome pro-
gram. Available online at: www.karobio.com/investormedia/
pressreleaser/pressr elease?pid= 639535 (accessed October 15,
2013).
324. Shiohara H, Nakamura T, Kikuchi N, Ozawa T, Nagano R,
Matsuzawa A, Ohnota H, Miyamoto T, Ichikawa K, Ha-
shizume K 2012 Discovery of novel indane derivatives as
INVESTIGATING THYROID HORMONE ECONOMY AND ACTION 153
liver-selective thyroid hormone receptor beta (TRbeta) ag-
onists for the treatment of dyslipidemia. Bioorg Med Chem
20:3622–3634.
325. Ito BR, Zhang BH, Cable EE, Song X, Fujitaki JM, MacK-
enna DA, Wilker CE, Chi B, van Poelje PD, Linemeyer DL,
Erion MD 2009 Thyroid hormone beta receptor activa-
tion has additive cholesterol lowering activity in combi-
nation with atorvastatin in rabbits, dogs and monkeys. Br J
Pharmacol 156:454–465.
326. Cable EE, Finn PD, Stebbins JW, Hou J, Ito BR, van Poelje
PD, Linemeyer DL, Erion MD 2009 Reduction of hepatic
steatosis in rats and mice after treatment with a liver-
targeted thyroid hormone receptor agonist. Hepatology
49:407–417.
327. Ocasio CA, Scanlan TS 2006 Design and characterization of
a thyroid hormone receptor alpha (TRalpha)-specific ago-
nist. ACS Chem Biol 1:585–593.
328. Denver RJ, Hu F, Scanlan TS, Furlow JD 2009 Thyroid hor-
mone receptor subtype specificity for hormone-dependent
neurogenesis in Xenopus laevis. Dev Biol 326:155–168.
329. Grijota-Martinez C, Samarut E, Scanlan TS, Morte B, Bernal
J 2011 In vivo activity of the thyroid hormone receptor
beta- and alpha-selective agonists GC-24 and CO
2
3 on rat
liver, heart, and brain. Endocrinology 152:1136–1142.
330. Morreale de Escobar G, Obrego
´
n MJ, Escobar del Rey F
2007 Iodine deficiency and brain development in the first
half of pregnancy. Public Health Nutr 10:1554–1570.
331. Schroder-van der Elst JP, van der Heide D, Kastelijn J,
Rousset B, Obregon MJ 2001 The expression of the sodium/
iodide symporter is up-regulated in the thyroid of fetuses
of iodine-deficient rats. Endocrinology 142:3736–3741.
332. Riesco G, Taurog A, Larsen PR 1976 Variations in the
response of the thyroid gland of the rat to different
low-iodine diets: correlation with iodine content of diet.
Endocrinology 99:270–280.
333. Riesco G, Taurog A, Larsen PR, Kru
¨
lich L 1977 Acute and
chronic responses to iodine deficiency in rats. Endo-
crinology 100:303–313.
334. Pazos-Moura CC, Moura EG, Dorris MM, Rehnmark S,
Melendez L, Silva JE, Taurog A 1991 Effect of iodine defi-
ciency and cold exposure on thyroxine 5¢-deiodinase ac-
tivity in various rat tissues. Am J Physiol 260:E175–E182.
335. Lavado-Autric R, Calvo RM, Martinez de Mena R, Mor-
reale de Escobar G, Obrego
´
n MJ 2013 Deiodinase activities
in thyroids and tissues of iodine deficient female rats. En-
docrinology 154:529–536.
336. Santisteban P, Obregon MJ, Rodriguez-Pena A, Lamas L,
Escobar del Rey F, Morreale de Escobar G 1982 Are iodine-
deficient rats euthyroid? Endocrinology 110:1780–1789.
337. Obregon MJ, Santisteban P, Rodriguez-Pena A, Pascual A,
Cartagena P, Ruiz-Marcos A, Lamas L, Escobar del Rey F,
Morreale de Escobar G 1984 Cerebral hypothyroidism in
rats with adult-onset iodine deficiency. Endocrinology
115:614–624.
338. Obregon MJ, Escobar del Rey F, Morreale de Escobar G
2005 The effects of iodine deficiency on thyroid hormone
deiodination. Thyroid 15:917–929.
339. Escobar del Rey F, Ruiz de Ona C, Bernal J, Obregon MJ,
Morreale de Escobar G 1989 Generalized deficiency of
3,5,3¢-triiodo-L-thyronine (T3) in tissues from rats on a low
iodine intake, despite normal circulating T
3
levels. Acta
Endocrinol (Copenh) 120:490–498.
340. Escobar del Rey F, Pastor R, Mallol J, Morreale de Escobar
G 1986 Effects of maternal iodine deficiency on the L-
thyroxine and 3,5,3¢-triiodo-L-thyronine contents of rat
embryonic tissues before and after onset of fetal thyroid
function. Endocrinology 118:1259–1265.
341. Obregon MJ, Ruiz de Ona C, Calvo R, Escobar del Rey F,
Morreale de Escobar G 1991 Outer ring iodothyronine
deiodinases and thyroid hormone economy: responses to
iodine deficiency in the rat fetus and neonate. En-
docrinology 129:2663–2673.
342. Martinez-Galan JR, Pedraza P, Santacana M, Escobar del
Rey F, Morreale de Escobar G, Ruiz-Marcos A 1997 Early
effects of iodine deficiency on radial glial cells of the hip-
pocampus of the rat fetus. A model of neurological cre-
tinism. J Clin Invest 99:2701–2709.
343. Martinez-Galan JR, Pedraza P, Santacana M, Escobar del
Rey F, Morreale de Escobar G, Ruiz-Marcos A 1997 Myelin
basic protein immunoreactivity in the internal capsule of
neonates from rats on a low iodine intake or on methyl-
mercaptoimidazole (MMI). Brain Res Dev Brain Res 101:
249–256.
344. Lavado-Autric R, Auso E, Garcia-Velasco JV, Arufe Mdel
C, Escobar del Rey F, Berbel P, Morreale de Escobar G 2003
Early maternal hypothyroxinemia alters histogenesis and
cerebral cortex cytoarchitecture of the progeny. J Clin In-
vest 111:1073–1082.
345. Morreale de Escobar G, Obregon MJ, Ruiz de On
˜
a C, Es-
cobar del Rey F 1988 Transfer of thyroxine from the mother
to the rat fetus near term: effects on brain 3,5,3¢-triiodo-
thyronine deficiency. Endocrinology 122:1521–1531.
346. Calvo R, Obregon MJ, Ruiz de Ona C, Escobar del Rey F,
Morreale de Escobar G 1990 Congenital hypothyroidism as
studied in rats: crucial role of maternal thyroxine (T4), but
not of 3,5,3¢ triiodothyronine (T3) in the protection of the
fetal brain. J Clin Invest 86:889–899.
347. Morreale de Escobar G, Calvo R, Obregon MJ, Escobar del
Rey F 1990 Contribution of maternal thyroxine to fetal
thyroxine pools in normal rats near term. Endocrinology
126:2765–2767.
348. Boelen A, Kwakkel J, Fliers E 2011 Beyond low plasma T3:
local thyroid hormone metabolism during inflammation
and infection. Endocr Rev 32:670–693.
349. Rocchi R, Kimura H, Tzou SC, Suzuki K, Rose NR, Pinch-
era A, Ladenson PW, Caturegli P 2007 Toll-like receptor-
MyD88 and Fc receptor pathways of mast cells mediate the
thyroid dysfunctions observed during nonthyroidal illness.
Proc Natl Acad Sci USA 104:6019–6024.
350. Mebis L, van den Berghe G 2009 The hypothalamus-pituitary-
thyroid axis in critical illness. Neth J Med 67:332–340.
351. Vella KR, Ramadoss P, Lam FS, Harris JC, Ye FD, Same PD,
O’Neill NF, Maratos-Flier E, Hollenberg AN 2011 NPY and
MC4R signaling regulate thyroid hormone levels during
fasting through both central and peripheral pathways. Cell
Metab 14:780–790.
352. Ahima RS, Prabakaran D, Mantzoros C, Qu D, Lowell B,
Maratos-Flier E, Flier JS 1996 Role of leptin in the neuro-
endocrine response to fasting. Nature 382:250–252.
353. Peeters RP, Wouters PJ, Kaptein E, van Toor H, Visser TJ,
Van den Berghe G 2003 Reduced activation and increased
inactivation of thyroid hormone in tissues of critically ill
patients. J Clin Endocrinol Metab 88:3202–3211.
354. Huang SA, Bianco AC 2008 Reawakened interest in type III
iodothyronine deiodinase in critical illness and injury. Nat
Clin Pract Endocrinol Metab 4:148–155.
355. Mebis L, Debaveye Y, Visser TJ, van den Berghe G 2006
Changes within the thyr oid axis during the course of
154 BIANCO ET AL.
critical illness. Endocrinol Metab Clin North Am 35:807–
821, x.
356. Goodman MN, Larsen PR, Kaplan MM, Aoki TT, Young
VR, Ruderman NB 1980 Starvation in the rat. II. Effect of
age and obesity on protein sparing and fuel metabolism.
Am J Physiol 239:E277–E286.
357. Wu SY 1990 The effect of fasting on thyroidal T4-5¢
monodeiodinating activity in mice. Acta Endocrinol (Co-
penhagen) 122:175–180.
358. Boelen A, van Beeren M, Vos X, Surovtseva O, Belegri E,
Saaltink DJ, Vreugdenhil E, Kalsbeek A, Kwakkel J, Fliers E
2012 Leptin administration restores the fasting-induced
increase of hepatic type 3 deiodinase expression in mice.
Thyroid 22:192–199.
359. Harris ARC, Fang SL, Vagenakis AG, Braverman LE 1978
Effect of starvation, nutrient replacement, and hypothy-
roidism on in vitro hepatic T
4
to T
3
conversion in the rat.
Metabolism 27:1680–1690.
360. Kinlaw WB, Schwartz HL, Oppenheimer JH 1985 De-
creased serum triiodothyronine in starving rats is due pri-
marily to diminished thyroidal secretion of thyroxine. J
Clin Invest 75:1238–1241.
361. Boelen A, Kwakkel J, Wiersinga WM, Fliers E 2006 Chronic
local inflammation in mice results in decreased TRH and
type 3 deiodinase mRNA expression in the hypothalamic
paraventricular nucleus independently of diminished food
intake. J Endocrinol 191:707–714.
362. Boelen A, Kwakkel J, Alkemade A, Renckens R, Kaptein E,
Kuiper G, Wiersinga WM, Visser TJ 2005 Induction of type
3 deiodinase activity in inflammatory cells of mice with
chronic local inflammation. Endocrinology 146:5128–5134.
363. Boelen A, Platvoet-ter Schiphorst MC, Wiersinga WM 1997
Immunoneutralization of interleukin-1, tumor necrosis
factor, interleukin-6 or interferon does not prevent the LPS-
induced sick euthyroid syndrome in mice. J Endocrinol
153:115–122.
364. Boelen A, Maas MA, Lowik CW, Platvoet MC, Wiersinga
WM 1996 Induced illness in interleukin-6 (IL-6) knock-out
mice: a causal role of IL-6 in the development of the low
3,5,3¢-triiodothyronine syndrome. Endocrinology 137:5250–
5254.
365. Boelen A, Platvoet-ter Schiphorst MC, Bakker O, Wiersinga
WM 1995 The role of cytokines in the lipopolysaccharide-
induced sick euthyroid syndrome in mice. J Endocrinol
146:475–483.
366. Bianco AC, Nunes MT, Hell NS, Maciel RM 1987 The role
of glucocorticoids in the stress-induced reduction of extra-
thyroidal 3,5,3¢-triiodothyronine generation in rats. En-
docrinology 120:1033–1038.
367. Mebis L, Debaveye Y, Ellger B, Derde S, Ververs EJ, Lan-
gouche L, Darras VM, Fliers E, Visser TJ, van den Berghe G
2009 Changes in the central component of the hypothala-
mus-pituitary-thyroid axis in a rabbit model of prolonged
critical illness. Crit Care 13:R147.
368. Weekers F, Van Herck E, Coopmans W, Michalaki M, Bowers
CY, Veldhuis JD, Van den Berghe G 2002 A novel in vivo
rabbit model of hypercatabolic critical illness reveals a biphasic
neuroendocrine stress response. Endocrinology 143:764–774.
369. Debaveye Y, Ellger B, Mebis L, Darras VM, Van den Berghe
G 2008 Regulation of tissue iodothyronine deiodinase ac-
tivity in a model of prolonged critical illness. Thyroid
18:551–560.
370. Boelen A, Kwakkel J, Wieland CW, St Germain DL, Fliers
E, Hernandez A 2009 Impaired bacterial clearance in type 3
deiodinase-deficient mice infected with Streptococcus
pneumoniae. Endocrinology 150:1984–1990.
371. Kondo K, Harbuz MS, Levy A, Lightman SL 1997 Inhibi-
tion of the hypothalamic-pituitary-thyroid axis in response
to lipopolysaccharide is independent of changes in circu-
lating corticosteroids. Neuroimmunomodulation 4:188–
194.
372. Fekete C, Sarkar S, Christoffolete MA, Emerson CH, Bianco
AC, Lechan RM 2005 Bacterial lipopolysaccharide (LPS)-
induced type 2 iodothyronine deiodinase (D2) activation in
the mediobasal hypothalamus (MBH) is independent of the
LPS-induced fall in serum thyroid hormone levels. Brain
Research 1056:97–99.
373. Fekete C, Gereben B, Doleschall M, Harney JW, Dora JM,
Bianco AC, Sarkar S, Liposits Z, Rand W, Emerson C,
Kacskovics I, Larsen PR, Lechan RM 2004 Lipopoly-
saccharide induces type 2 iodothyronine deiodinase in the
mediobasal hypothalamus: implications for the non-
thyroidal illness syndrome. Endocrinology 145:1649–1655.
374. Boelen A, Kwakkel J, Thijssen-Timmer DC, Alkemade A,
Fliers E, Wiersinga WM 2004 Simultaneous changes in
central and peripheral components of the hypothalamus-
pituitary-thyroid axis in lipopolysaccharide-induced acute
illness in mice. J Endocrinol 182:315–323.
375. Zeo
¨
ld A, Doleschall M, Haffner MC, Capelo LP, Menyhert
J, Liposits Z, da Silva WS, Bianco AC, Kacskovics I, Fekete
C, Gereben B 2006 Characterization of the nuclear factor-
kappa B responsiveness of the human DIO2 gene. En-
docrinology 147:4419–4429.
376. Lamirand A, Ramauge M, Pierre M, Courtin F 2011 Bac-
terial lipopolysaccharide induces type 2 deiodinase in cul-
tured rat astrocytes. J Endocrinol 208:183–192.
377. Sap J, Mun
˜
oz A, Damm K, Goldberg Y, Ghysdael J, Leutz
A, Beug H, Vennstro
¨
m B 1986 The c-erb-A protein is a high-
affinity receptor for thyroid hormone. Nature 324:635–640.
378. Weinberger C, Thompson CC, Ong ES, Lebo R, Gruol DJ,
Evans RM 1986 The c-erb-A gene encodes a thyroid hor-
mone receptor. Nature 324:641–646.
379. Brent GA, Moore DD, Larsen PR 1991 Thyroid hormone
regulation of gene expression. Annu Rev Physiol 53:17–35.
380. Glass CK, Franco R, Weinberger C, Albert VR, Evans RM,
Rosenfeld MG 1987 A c-erb-A binding site in rat growth
hormone gene mediates trans-activation by thyroid hor-
mone. Nature 329:738–741.
381. Sap J, Mun
˜
oz A, Schmitt J, Stunnenberg H, Vennstro
¨
mB
1989 Repression of transcription mediated at a thyroid
hormone response element by the v-erb-A oncogene prod-
uct. Nature 340:242–244.
382. Cheng SY 2000 Multiple mechanisms for regulation of the
transcriptional activity of thyroid hormone receptors. Rev
Endocr Metab Disord 1:9–18.
383. Flamant F, Gauthier K, Samarut J 2007 Thyroid hormones
signaling is getting more complex: STORMs are coming.
Mol Endocrinol 21:321–333.
384. [Deleted.]
385. Chan IH, Privalsky ML 2009 Isoform-specific transcrip-
tional activity of overlapping target genes that respond to
thyroid hormone receptors alpha1 and beta1. Mol En-
docrinol 23:1758–1775.
386. Gloss B, Trost S, Bluhm W, Swanson E, Clark R, Winkfein
R, Janzen K, Giles W, Chassande O, Samarut J, Dillmann W
2001 Cardiac ion channel expression and contractile func-
tion in mice with deletion of thyroid hormone receptor al-
pha or beta. Endocrinology 142:544–550.
INVESTIGATING THYROID HORMONE ECONOMY AND ACTION 155
387. Wan W, Farboud B, Privalsky ML 2005 Pituitary resistance
to thyroid hormone syndrome is associated with T
3
re-
ceptor mutants that selectively impair beta2 isoform func-
tion. Mol Endocrinol 19:1529–1542.
388. Yang Z, Privalsky ML 2001 Isoform-specific transcriptional
regulation by thyroid hormone receptors: hormone-
independent activation operates through a steroid receptor
mode of co-activator interaction. Mol Endocrinol 15:1170–
1185.
389. Zhu XG, McPhie P, Lin KH, Cheng SY 1997 The differential
hormone-dependent transcriptional activation of thyroid
hormone receptor isoforms is mediated by interplay of
their domains. J Biol Chem 272:9048–9054.
390. Flores-Morales A, Gullberg H, Fernandez L, Stahlberg N,
Lee NH, Vennstrom B, Norstedt G 2002 Patterns of liver
gene expression governed by TRbeta. Mol Endocrinol
16:1257–1268.
391. Munoz A, Rodriguez-Pena A, Perez-Castillo A, Ferreiro B,
Sutcliffe JG, Bernal J 1991 Effects of neonatal hypothy-
roidism on rat brain gene expression. Mol Endocrinol
5:273–280.
392. Thompson CC 1996 Thyroid hormone-responsive genes in
developing cerebellum include a novel synaptotagmin and
a hairless homolog. J Neurosci 16:7832–7840.
393. Yen PM, Feng X, Flamant F, Chen Y, Walker RL, Weiss RE,
Chassande O, Samarut J, Refetoff S, Meltzer PS 2003 Effects
of ligand and thyroid hormone receptor isoforms on he-
patic gene expression profiles of thyroid hormone receptor
knockout mice. EMBO Rep 4:581–587.
394. Bigler J, Eisenman RN 1994 Isolation of a thyroid hormone-
responsive gene by immunoprecipitation of thyroid hor-
mone receptor-DNA complexes. Mol Cell Biol 14:7621–7632.
395. Royland JE, Parker JS, Gilbert ME 2008 A genomic analysis
of subclinical hypothyroidism in hippocampus and neo-
cortex of the developing rat brain. J Neuroendocrinol
20:1319–1338.
396. Guadan
˜
o-Ferraz A, Esca
´
mez MJ, Morte B, Vargiu P, Bernal
J 1997 Transcriptional induction of RC3/neurogranin by
thyroid hormone: differential neuronal sensitivity is not
correlated with thyroid hormone receptor distribution in
the brain. Brain Res Mol Brain Res 49:37–44.
397. Anderson GW, Larson RJ, Oas DR, Sandhofer CR, Schwartz
HL, Mariash CN, Oppenheimer JH 1998 Chicken ovalbumin
upstream promoter-transcription factor (COUP-TF) modulates
expression of the Purkinje cell protein-2 gene. A potential role
for COUP-TF in repressing premature thyroid hormone action
in the developing brain. J Biol Chem 273:16391–16399.
398. Ng L, Lu A, Swaroop A, Sharlin DS, Swaroop A, Forrest D
2011 Two transcription factors can direct three photore-
ceptor outcomes from rod precursor cells in mouse retinal
development. J Neurosci 31:11118–11125.
399. [Deleted.]
400. Dong H, Yauk CL, Rowan-Carroll A, You SH, Zoeller RT,
Lambert I, Wade MG 2009 Identification of thyroid hormone
receptor binding sites and target genes using ChIP-on-chip
in developing mouse cerebellum. PLoS One 4:e4610.
401. Morte B, Diez D, Auso E, Belinchon MM, Gil-Ibanez P,
Grijota-Martinez C, Navarro D, Morreale de Escobar G,
Berbel P, Bernal J 2010 Thyroid hormone regulation of gene
expression in the developing rat fetal cerebral cortex: pro-
minent role of the Ca2 + /calmodulin-dependent protein
kinase IV pathway. Endocrinology 151:810–820.
402. Kahaly GJ, Dillmann WH 2005 Thyroid hormone action in
the heart. Endocr Rev 26:704–728.
403. Ribeiro MO, Bianco SD, Kaneshige M, Schultz JJ, Cheng SY,
Bianco AC, Brent GA 2010 Expression of uncoupling pro-
tein 1 in mouse brown adipose tissue is thyroid hormone
receptor-beta isoform specific and required for adaptive
thermogenesis. Endocrinology 151:432–440.
404. Wang AM, Doyle MV, Mark DF 1989 Quantitation of
mRNA by the polymerase chain reaction. Proc Natl Acad
Sci USA 86:9717–9721.
405. Sood A, Schwartz HL, Oppenheimer JH 1996 Tissue-
specific regulation of malic enzyme by thyroid hormone in
the neonatal rat. Biochem Biophys Res Commun 222:287–291.
406. Bustin SA 2000 Absolute quantification of mRNA using
real-time reverse transcription polymerase chain reaction
assays. J Mol Endocrinol 25:169–193.
407. Fraga D, Meulia T, Fenster S 2008 Real-time PCR. In: Ca-
valcanti ARO, Stover N (eds) Current Protocols Essential
Laboratory Techniques. John Wiley and Sons, New York,
pp 10.3.1–10.3.33.
408. Nolan T, Hands RE, Bustin SA 2006 Quantification of
mRNA using real-time RT-PCR. Nat Protoc 1:1559–1582.
409. Ness GC, Pendleton LC 1991 Thyroid hormone increases
glyceraldehyde 3-phosphate dehydrogenase gene expres-
sion in rat liver. FEBS Lett 288:21–22.
410. Poddar R, Paul S, Chaudhury S, Sarkar PK 1996 Regulation of
actin and tubulin gene expression by thyroid hormone during
rat brain development. Brain Res Mol Brain Res 35:111–118.
411. Visser WE, Heemstra KA, Swagemakers SM, Ozgur Z,
Corssmit EP, Burggraaf J, van Ijcken WF, van der Spek PJ,
Smit JW, Visser TJ 2009 Physiological thyroid hormone
levels regulate numerous skeletal muscle transcripts. J Clin
Endocrinol Metab 94:3487–3496.
412. Diez D, Grijota-Martinez C, Agretti P, De Marco G, To-
nacchera M, Pinchera A, Morreale de Escobar G, Bernal J,
Morte B 2008 Thyroid hormone action in the adult brain:
gene expression profiling of the effects of single and mul-
tiple doses of triiodo-L-thyronine in the rat striatum. En-
docrinology 149:3989–4000.
413. Chiappini F, Ramadoss P, Vella KR, Cunha LL, Ye FD,
Stuart RC, Nillni EA, Hollenberg AN 2013 Family members
CREB and CREM control thyrotropin-releasing hormone
(TRH) expression in the hypothalamus. Mol Cell Endo-
crinol 365:84–94.
414. Livak KJ, Schmittgen TD 2001 Analysis of relative gene
expression data using real-time quantitative PCR and the
2(-Delta Delta C(T)) method. Methods 25:402–408.
415. Nolan T, Hands RE, Bustin SA 2006 Quantification of
mRNA using real-time RT-PCR. Nat Protoc 1:1559–1582.
416. [Deleted.]
417. Grewal A, Lambert P, Stockton J 2007 Analysis of expres-
sion data: an overview. Curr Protoc Bioinformatics Chap-
ter 7:Unit 7.1.
418. Hawkins RD, Hon GC, Ren B 2010 Next-generation geno-
mics: an integrative approach. Nat Rev Genet 11:476–486.
419. Espina V, Milia J, Wu G, Cowherd S, Liotta LA 2006 Laser
capture microdissection. Methods Mol Biol 319:213–229.
420. Galbraith DW, Elumalai R, Gong FC 2004 Integrative flow
cytometric and microarray approaches for use in tran-
scriptional profiling. Methods Mol Biol 263:259–280.
421. Schoonover CM, Seibel MM, Jolson DM, Stack MJ, Rahman
RJ, Jones SA, Mariash CN, Anderson GW 2004 Thyroid hor-
mone regulates oligodendrocyte accumulation in developing
rat brain white matter tracts. Endocrinology 145:5013–5020.
422. Sharlin DS, Tighe D, Gilbert ME, Zoeller RT 2008 The
balance between oligodendrocyte and astrocyte production
156 BIANCO ET AL.
in major white matter tracts is linearly related to serum
total thyroxine. Endocrinology 149:2527–2536.
423. Phan TQ, Jow MM, Privalsky ML 2010 DNA recognition by
thyroid hormone and retinoic acid receptors: 3,4,5 rule
modified. Mol Cell Endocrinol 319:88–98.
424. Chen JD, Evans RM 1995 A transcriptional co-repressor
that interacts with nuclear hormone receptors. Nature
377:454–457.
425. Fondell JD, Ge H, Roeder RG 1996 Ligand induction of a
transcriptionally active thyroid hormone receptor coacti-
vator complex. Proc Natl Acad Sci USA 93:8329–8333.
426. Ho
¨
rlein AJ, Na
¨
a
¨
r AM, Heinzel T, Torchia J, Gloss B, Kur-
okawa R, Ryan A, Kamei Y, So
¨
derstro
¨
m M, Glass CK,
Rosenfeld MG 1995 Ligand-independent repression by the
thyroid hormone receptor mediated by a nuclear receptor
co-repressor. Nature 377:397–404.
427. Gould SJ, Subramani S 1988 Firefly luciferase as a tool in
molecular and cell biology. Anal Biochem 175:5–13.
428. Chan IH, Borowsky AD, Privalsky ML 2008 A cautionary
note as to the use of pBi-L and related luciferase/transgenic
vectors in the study of thyroid endocrinology. Thyroid
18:665–666.
429. Shifera AS, Hardin JA 2010 Factors modulating expression
of Renilla luciferase from control plasmids used in lucifer-
ase reporter gene assays. Anal Biochem 396:167–172.
430. Jones I, Ng L, Liu H, Forrest D 2007 An intron control
region differentially regulates expression of thyroid hor-
mone receptor beta2 in the cochlea, pituitary, and cone
photoreceptors. Mol Endocrinol 21:1108–1119.
431. Sjo
¨
berg M, Vennstro
¨
m B, Forrest D 1992 Thyroid hormone
receptors in chick retinal development: differential expres-
sion of mRNAs for alpha and N-terminal variant beta re-
ceptors. Development 114:39–47.
432. Wood WM, Dowding JM, Haugen BR, Bright TM, Gordon
DF, Ridgway EC 1994 Structural and functional charac-
terization of the genomic locus encoding the murine beta 2
thyroid hormone receptor. Mol Endocrinol 8:1605–1617.
433. Sap J, de Magistris L, Stunnenberg H, Vennstrom B 1990 A
major thyroid hormone response element in the third in-
tron of the rat growth hormone gene. EMBO J 9:887–896.
434. Umesono K, Murakami KK, Thompson CC, Evans RM
1991 Direct repeats as selective response elements for the
thyroid hormone, retinoic acid, and vitamin D3 receptors.
Cell 65:1255–1266.
435. Hellman LM, Fried MG 2007 Electrophoretic mobility shift
assay (EMSA) for detecting protein-nucleic acid interac-
tions. Nat Protoc 2:1849–1861.
436. Ng L, Forrest D, Haugen BR, Wood WM, Curran T 1995 N-
terminal variants of thyroid hormone receptor beta: dif-
ferential function and potential contribution to syndrome
of resistance to thyroid hormone. Mol Endocrinol 9:1202–
1213.
437. Yen PM, Darling DS, Carter RL, Forgione M, Umeda PK,
Chin WW 1992 Triiodothyronine (T3) decreases binding
to DNA by T3-receptor homodimers but not receptor-
auxiliary protein heterodimers. J Biol Chem 267:3565–3568.
438. Belakavadi M, Saunders J, Weisleder N, Raghava PS,
Fondell JD 2010 Repression of cardiac phospholamban
gene expression is mediated by thyroid hormone receptor-
{alpha}1 and involves targeted covalent histone modifica-
tions. Endocrinology 151:2946–2956.
439. Chiamolera MI, Sidhaye AR, Matsumoto S, He Q, Ha-
shimoto K, Ortiga-Carvalho TM, Wondisford FE 2012
Fundamentally distinct roles of thyroid hormone receptor
isoforms in a thyrotroph cell line are due to differential
DNA binding. Mol Endocrinol 26:926–939.
440. Bilesimo P, Jolivet P, Alfama G, Buisine N, Le Mevel S,
Havis E, Demeneix BA, Sachs LM 2011 Specific histone
lysine 4 methylation patterns define TR-binding capacity
and differentiate direct T
3
responses. Mol Endocrinol
25:225–237.
441. Matsuura K, Fujimoto K, Fu L, Shi YB 2012 Liganded
thyroid hormone receptor induces nucleosome removal
and histone modifications to activate transcription during
larval intestinal cell death and adult stem cell development.
Endocrinology 153:961–972.
442. Havis E, Le Mevel S, Morvan Dubois G, Shi DL, Scanlan
TS, Demeneix BA, Sachs LM 2006 Unliganded thyroid
hormone receptor is essential for Xenopus laevis eye de-
velopment. EMBO J 25: 4943–4951.
443. Carey MF, Peterson CL, Smale ST 2009 Chromatin immu-
noprecipitation (ChIP). Cold Spring Harb Protoc 2009:pdb.
prot5279.
444. Wagschal A, Delaval K, Pannetier M, Arnaud P, Feil R 2007
PCR-based analysis of immunoprecipitated chromatin.
Cold Spring Harb Protoc 2007:pdb.prot4768.
445. Morte B, Manzano J, Scanlan T, Vennstro
¨
m B, Bernal J 2002
Deletion of the thyroid hormone receptor alpha 1 prevents
the structural alterations of the cerebellum induced by
hypothyroidism. Proc Natl Acad Sci USA 99:3985–3989.
446. Abel ED, Ahima RS, Boers ME, Elmquist JK, Wondisford
FE 2001 Critical role for thyroid hormone receptor beta2 in
the regulation of paraventricular thyrotropin-releasing
hormone neurons. J Clin Invest 107:1017–1023.
447. Forrest D, Hanebuth E, Smeyne RJ, Everds N, Stewart CL,
Wehner JM, Curran T 1996 Recessive resistance to thyroid
hormone in mice lacking thyroid hormone receptor beta:
evidence for tissue-specific modulation of receptor func-
tion. EMBO J 15:3006–3015.
448. Forrest D, Erway LC, Ng L, Altschuler R, Curran T 1996
Thyroid hormone receptor beta is essential for develop-
ment of auditory function. Nat Genet 13:354–357.
449. Ng L, Hurley JB, Dierks B, Srinivas M, Salto C, Vennstrom
B, Reh TA, Forrest D 2001 A thyroid hormone receptor that
is required for the development of green cone photore-
ceptors. Nat Genet 27:94–98.
450. Amma LL, Campos-Barros A, Wang Z, Vennstrom B,
Forrest D 2001 Distinct tissue-specific roles for thyroid
hormone receptors beta and alpha1 in regulation of type 1
deiodinase expression. Mol Endocrinol 15:467–475.
451. Wikstro
¨
m L, Johansson C, Salto
´
C, Barlow C, Campos
Barros A, Baas F, Forrest D, Thore
´
n P, Vennstro
¨
m B 1998
Abnormal heart rate and body temperature in mice lacking
thyroid hormone receptor alpha 1. EMBO J 17:455–461.
452. Fraichard A, Chassande O, Plateroti M, Roux JP, Trouillas J,
Dehay C, Legrand C, Gauthier K, Kedinger M, Malaval L,
Rousset B, Samarut J 1997 The T3R alpha gene encoding a
thyroid hormone receptor is essential for post-natal devel-
opment and thyroid hormone production. EMBO J 16:
4412–4420.
453. Bassett JH, Williams GR 2009 The skeletal phenotypes of
TRalpha and TRbeta mutant mice. J Mol Endocrinol 42:
269–282.
454. Guadan
˜
o-Ferraz A, Benavides-Piccione R, Venero C, Lan-
cha C, Vennstro
¨
m B, Sandi C, DeFelipe J, Bernal J 2003 Lack
of thyroid hormone receptor alpha1 is associated with se-
lective alterations in behavior and hippocampal circuits.
Mol Psychiatry 8:30–38.
INVESTIGATING THYROID HORMONE ECONOMY AND ACTION 157
455. [Deleted.]
456. Vennstro
¨
m B, Mittag J, Wallis K 2008 Severe psychomotor
and metabolic damages caused by a mutant thyroid hor-
mone receptor alpha 1 in mice: can patients with a similar
mutation be found and treated? Acta Paediatr 97:1605–
1610.
457. Gauthier K, Chassande O, Plateroti M, Roux JP, Legrand C,
Pain B, Rousset B, Weiss R, Trouillas J, Samarut J 1999 Dif-
ferent functions for the thyroid hormone receptors TRalpha
and TRbeta in the control of thyroid hormone production
and post-natal development. EMBO J 18:623–631.
458. Refetoff S 2003 Resistance to thyroid hormone with and
without receptor gene mutations. Annales d’endocrinologie
64:23–25.
459. Bochukova E, Schoenmakers N, Agostini M, Schoenmakers
E, Rajanayagam O, Keogh JM, Henning E, Reinemund J,
Gevers E, Sarri M, Downes K, Offiah A, Albanese A, Hal-
sall D, Schwabe JW, Bain M, Lindley K, Muntoni F, Vargha-
Khadem F, Dattani M, Farooqi IS, Gurnell M, Chatterjee K
2012 A mutation in the thyroid hormone receptor alpha
gene. N Engl J Med 366:243–249.
460. van Mullem A, van Heerebeek R, Chrysis D, Visser E,
Medici M, Andrikoula M, Tsatsoulis A, Peeters R, Visser TJ
2012 Clinical phenotype and mutant TRalpha1. N Engl J
Med 366:1451–1453.
461. Cheng SY 2005 Thyroid hormone receptor mutations and
disease: beyond thyroid hormone resistance. Trends En-
docrinol Metab 16:176–182.
462. Hashimoto K, Curty FH, Borges PP, Lee CE, Abel ED,
Elmquist JK, Cohen RN, Wondisford FE 2001 An un-
liganded thyroid hormone receptor causes severe neuro-
logical dysfunction. Proc Natl Acad Sci USA 98:3998–4003.
463. Kaneshige M, Kaneshige K, Zhu X, Dace A, Garrett L,
Carter TA, Kazlauskaite R, Pankratz DG, Wynshaw-Boris
A, Refetoff S, Weintraub B, Willingham MC, Barlow C,
Cheng S 2000 Mice with a targeted mutation in the thyroid
hormone beta receptor gene exhibit impaired growth and
resistance to thyroid hormone. Proc Natl Acad Sci USA
97:13209–13214.
464. Liu YY, Schultz JJ, Brent GA 2003 A thyroid hormone re-
ceptor alpha gene mutation (P398H) is associated with
visceral adiposity and impaired catecholamine-stimulated
lipolysis in mice. J Biol Chem 278:38913–38920.
465. Ortiga-Carvalho TM, Shibusawa N, Nikrodhanond A,
Oliveira KJ, Machado DS, Liao XH, Cohen RN, Refetoff S,
Wondisford FE 2005 Negative regulation by thyroid hor-
mone receptor requires an intact coactivator-binding sur-
face. J Clin Invest 115:2517–2523.
466. [Deleted.]
467. Quignodon L, Legrand C, Allioli N, Guadano-Ferraz A,
Bernal J, Samarut J, Flamant F 2004 Thyroid hormone sig-
naling is highly heterogeneous during pre- and postnatal
brain development. J Mol Endocrinol 33:467–476.
468. Nucera C, Muzzi P, Tiveron C, Farsetti A, La Regina F,
Foglio B, Shih SC, Moretti F, Della Pietra L, Mancini F,
Sacchi A, Trimarchi F, Vercelli A, Pontecorvi A 2010 Ma-
ternal thyroid hormones are transcriptionally active during
embryo-foetal development: results from a novel trans-
genic mouse model. J Cell Mol Med 14:2417–2435.
469. Wallis K, Dudazy S, van Hogerlinden M, Nordstrom K,
Mittag J, Vennstrom B 2010 The thyroid hormone receptor
alpha1 protein is expressed in embryonic postmitotic neu-
rons and persists in most adult neurons. Mol Endocrinol
24:1904–1916.
470. Williams GR 2008 Neurodevelopmental and neurophysio-
logical actions of thyroid hormone. J Neuroendocrinol
20:784–794.
471. Zoeller RT, Rovet J 2004 Timing of thyroid hormone action
in the developing brain: clinical observations and experi-
mental findings. J Neuroendocrinol 16:809–818.
472. Rogers DC, Fisher EM, Brown SD, Peters J, Hunter AJ,
Martin JE 1997 Behavioral and functional analysis of mouse
phenotype: SHIRPA, a proposed protocol for comprehen-
sive phenotype assessment. Mamm Genome 8:
711–713.
473. Brown SD, Chambon P, and de Angelis MH 2005 EM-
PReSS: standardized phenotype screens for functional an-
notation of the mouse genome. Nat Genet 37:1155.
474. Skarnes WC, Rosen B, West AP, Koutsourakis M, Bushell
W, Iyer V, Mujica AO, Thomas M, Harrow J, Cox T, Jack-
son D, Severin J, Biggs P, Fu J, Nefedov M, de Jong PJ,
Stewart AF, Bradley A 2011 A conditional knockout re-
source for the genome-wide study of mouse gene function.
Nature 474:337–342.
475. Wheeler SM, Willoughby KA, McAndrews MP, Rovet JF
2011 Hippocampal size and memory functioning in chil-
dren and adolescents with congenital hypothyroidism. J
Clin Endocrinol Metab 96: E1427–1434.
476. Morreale de Escobar G, Ruiz Marcos A, Escobar del Rey F
1983 Thyroid hormones and the developing brain. In:
Dussault JH, Walker P (eds) Congenital Hypothyroidism.
Marcel Dekker, New York, pp 85–126.
477. Castro MI, Alex S, Young RA, Braverman LE, Emerson CH
1986 Total and free serum thyroid hormone concentrations
in fetal and adult pregnant and nonpregnant guinea pigs.
Endocrinology 118:533–537.
478. Wallis K, Sjo
¨
gren M, van Hogerlinden M, Silberberg G,
Fisahn A, Nordstro
¨
m K, Larsson L, Westerblad H, Morreale
de Escobar G, Shupliakov O, Vennstro
¨
m B 2008 Locomotor
deficiencies and aberrant development of subtype-specific
GABAergic interneurons caused by an unliganded thyroid
hormone receptor alpha1. J Neurosci 28:1904–1915.
479. Fekete C, Freitas BC, Zeo
¨
ld A, Wittmann G, Ka
´
da
´
r A, Li-
posits Z, Christoffolete MA, Singru P, Lechan RM, Bianco
AC, Gereben B 2007 Expression patterns of WSB-1 and USP-
33 underlie cell-specific posttranslational control of type 2
deiodinase in the rat brain. Endocrinology 148:4865–4874.
480. Bernal J, Guadano-Ferraz A 2002 Analysis of thyroid
hormone-dependent genes in the brain by in situ hybrid-
ization. Methods Mol Biol 202:71–90.
481. Guadano-Ferraz A, Escamez MJ, Rausell E, Bernal J 1999
Expression of type 2 iodothyronine deiodinase in hypo-
thyroid rat brain indicates an important role of thyroid
hormone in the development of specific primary sensory
systems. J Neurosci 19:3430–3439.
482. Iniguez MA, De Lecea L, Guadano-Ferraz A, Morte B,
Gerendasy D, Sutcliffe JG, Bernal J 1996 Cell-specific effects
of thyroid hormone on RC3/neurogranin expression in rat
brain. Endocrinology 137:1032–1041.
483. Venero C, Guadano-Ferraz A, Herrero AI, Nordstrom K,
Manzano J, Morreale de Escobar G, Bernal J, Vennstrom B
2005 Anxiety, memory impairment, and locomotor dysfunc-
tion caused by a mutant thyroid hormone receptor alpha1 can
be ameliorated by T
3
treatment. Genes Dev 19:2152–2163.
484. Hrabovszky E, Petersen SL 2002 Increased concentrations of
radioisotopically-labeled complementary ribonucleic acid
probe, dextran sulfate, and dithiothreitol in the hybridization
buffer can improve results of in situ hybridization histo-
chemistry. J Histochem Cytochem 50:1389–1400.
158 BIANCO ET AL.
485. Campos-Barros A, Amma LL, Faris JS, Shailam R, Kelley
MW, Forrest D 2000 Type 2 iodothyronine deiodinase ex-
pression in the cochlea before the onset of hearing. Proc
Natl Acad Sci USA 97:1287–1292.
486. Lechan RM, Wu P, Jackson, IMD, Wolfe H, Cooperman S,
Mandel G, Goodman RH 1986 Thyrotropin-releasing hor-
mone precursor: characterization in rat brain. Science
231:159–161.
487. Segerson TP, Kauer J, Wolfe HC, Mobtaker H, Wu P,
Jackson IM, Lechan RM 1987 Thyroid hormone regulates
TRH biosynthesis in the paraventricular nucleus of the rat
hypothalamus. Science 238:78–80.
488. Dyess EM, Segerson TP, Liposits Z, Paull WK, Kaplan MM,
Wu P, Jackson, IMD, Lechan RM 1988 Triiodothyronine
exerts direct cell-specific regulation of thyrotropin-releasing
hormone gene expression in the hypothalamic para-
ventricular nucleus. Endocrinology 123:2291–2297.
489. Kakucska I, Rand W, Lechan RM 1992 Thyrotropin-
releasing hormone (TRH) gene expression in the hypotha-
lamic paraventricular nucleus is dependent upon feedback
regulation by both triiodothyronine and thyroxine. En-
docrinology 130:2845–2850.
490. [Deleted.]
491. Espina V, Wulfkuhle JD, Calvert VS, VanMeter A, Zhou W,
Coukos G, Geho DH, Petricoin EF 3rd, Liotta LA 2006
Laser-capture microdissection. Nat Protoc 1:586–603.
492. Hernandez A, Morte B, Belinchon MM, Ceballos A, Bernal J
2012 Critical role of types 2 and 3 deiodinases in the neg-
ative regulation of gene expression by t3 in the mouse ce-
rebral cortex. Endocrinology 153:2919–2928.
493. Dong H, Wade M, Williams A, Lee A, Douglas GR, Yauk C
2005 Molecular insight into the effects of hypothyroidism
on the developing cerebellum. Biochem Biophys Res
Commun 330:1182–1193.
494. Chatonnet F, Guyot R, Picou F, Bondesson M, Flamant F
2012 Genome-wide search reveals the existence of a limited
number of thyroid hormone receptor alpha target genes in
cerebellar neurons. PLoS One 7:e30703.
495. Silva JE, Rudas P 1990 Effect of congenital hypothyroidism
on microtubule-associated protein-2 expression in the cer-
ebellum of the rat. Endocrinology 126:1276–1282.
496. Sharlin DS, Visser TJ, Forrest D 2011 Developmental and
cell-specific expression of thyroid hormone transporters in
the mouse cochlea. Endocrinology 152:5053–5064.
497. Alvarez-Dolado M, Ruiz M, Del
´
o JA, Alca
´
ntara S, Bur-
gaya F, Sheldon M, Nakajima K, Bernal J, Howell BW,
Curran T, Soriano E, Mun
˜
oz A 1999 Thyroid hormone
regulates reelin and dab1 expression during brain devel-
opment. J Neurosci 19: 6979–6993.
498. Alvarez-Dolado M, Gonzalez-Sancho JM, Bernal J, Munoz
A 1998 Developmental expression of the tenascin-C is al-
tered by hypothyroidism in the rat brain. Neuroscience
84:309–322.
499. Cordas EA, Ng L, Hernandez A, Kaneshige M, Cheng SY,
Forrest D 2012 Thyroid hormone receptors control devel-
opmental maturation of the middle ear and the size of the
ossicular bones. Endocrinology 153:1548–1560.
500. [Deleted.]
501. Winter H, Ru
¨
ttiger L, Mu
¨
ller M, Kuhn S, Brandt N, Zim-
mermann U, Hirt B, Bress A, Sausbier M, Conscience A,
Flamant F, Tian Y, Zuo J, Pfister M, Ruth P, Lo
¨
wenheim H,
Samarut J, Engel J, Knipper M 2009 Deafness in TRbeta
mutants is caused by malformation of the tectorial mem-
brane. J Neurosci 29:
2581–2587.
502. Griffith AJ, Szymko YM, Kaneshige M, Quinonez RE, Ka-
neshige K, Heintz KA, Mastroianni MA, Kelley MW,
Cheng SY 2002 Knock-in mouse model for resistance to
thyroid hormone (RTH): an RTH mutation in the thyroid
hormone receptor beta gene disrupts cochlear morpho-
genesis. J Assoc Res Otolaryngol 3:279–288.
503. Applebury ML, Farhangfar F, Glosmann M, Hashimoto K,
Kage K, Robbins JT, Shibusawa N, Wondisford FE, Zhang
H 2007 Transient expression of thyroid hormone nuclear
receptor TRbeta2 sets S opsin patterning during cone
photoreceptor genesis. Dev Dyn 236:1203–1212.
504. Lu A, Ng L, Ma M, Kefas B, Davies TF, Hernandez A, Chan
CC, Forrest D 2009 Retarded developmental expression
and patterning of retinal cone opsins in hypothyroid mice.
Endocrinology 150:1536–1544.
505. [Deleted.]
506. Glaschke A, Glosmann M, Peichl L 2010 Developmental
changes of cone opsin expression but not retinal morphol-
ogy in the hypothyroid Pax8 knockout mouse. Invest
Ophthalmol Vis Sci 51:1719–1727.
507. Gilbert ME, Sui L, Walker MJ, Anderson W, Thomas S,
Smoller SN, Schon JP, Phani S, Goodman JH 2007 Thyroid
hormone insufficiency during brain development reduces
parvalbumin immunoreactivity and inhibitory function in
the hippocampus. Endocrinology 148:92–102.
508. Gilbert ME, Paczkowski C 2003 Propylthiouracil (PTU)-
induced hypothyroidism in the developing rat impairs
synaptic transmission and plasticity in the dentate gyrus
of the adult hippocampus. Brain Res Dev Brain Res 145:
19–29.
509. Sui L, Gilbert ME 2003 Pre- and postnatal propylthiouracil-
induced hypothyroidism impairs synaptic transmission
and plasticity in area CA1 of the neonatal rat hippocampus.
Endocrinology 144:4195–4203.
510. Gilbert ME 2004 Alterations in synaptic transmission and
plasticity in area CA1 of adult hippocampus following
developmental hypothyroidism. Brain Res Dev Brain Res
148:11–18.
511. Pilhatsch M, Winter C, Nordstrom K, Vennstrom B, Bauer
M, Juckel G 2010 Increased depressive behaviour in mice
harboring the mutant thyroid hormone receptor alpha 1.
Behav Brain Res 214:187–192.
512. Mombereau C, Kaupmann K, Gassmann M, Bettler B, van
der Putten H, Cryan JF 2005 Altered anxiety and depres-
sion-related behaviour in mice lacking GABAB(2) receptor
subunits. Neuroreport 16:307–310.
513. Reif A, Schmitt A, Fritzen S, Chourbaji S, Bartsch C, Urani
A, Wycislo M, Mo
¨
ssner R, Sommer C, Gass P, Lesch KP
2004 Differential effect of endothelial nitric oxide synthase
(NOS-III) on the regulation of adult neurogenesis and be-
haviour. Eur J Neurosci 20:885–895.
514. Lo
´
pez M, Varela L, Va
´
zquez MJ, Rodrı
´
guez-Cuenca S,
Gonza
´
lez CR, Velagapudi VR, Morgan DA, Schoenmakers
E, Agassandian K, Lage R, Martı
´
nez de Morentin PB, Tovar
S, Nogueiras R, Carling D, Lelliott C, Gallego R, Oresic M,
Chatterjee K, Saha AK, Rahmouni K, Die
´
guez C, Vidal-
Puig A 2010 Hypothalamic AMPK and fatty acid metabo-
lism mediate thyroid regulation of energy balance. Nat
Med 16:1001–1008.
515. Coppola A, Liu ZW, Andrews ZB, Paradis E, Roy MC,
Friedman JM, Ricquier D, Richard D, Horvath TL, Gao XB,
Diano S 2007 A central thermogenic-like mechanism in
feeding regulation: an interplay between arcuate nucleus T
3
and UCP2. Cell Metab 5:21–33.
INVESTIGATING THYROID HORMONE ECONOMY AND ACTION 159
516. Kong WM, Martin NM, Smith KL, Gardiner JV, Connoley
IP, Stephens DA, Dhillo WS, Ghatei MA, Small CJ, Bloom
SR 2004 Triiodothyronine stimulates food intake via the
hypothalamic ventromedial nucleus independent of
changes in energy expenditure. Endocrinology 14 5:5 252 –
5258.
517. Barrett P, Ebling FJ, Schuhler S, Wilson D, Ross AW,
Warner A, Jethwa P, Boelen A, Visser TJ, Ozanne DM,
Archer ZA, Mercer JG, Morgan PJ 2007 Hypothalamic
thyroid hormone catabolism acts as a gatekeeper for the
seasonal control of body weight and reproduction. En-
docrinology 148:3608–3617.
518. Fekete C, Legradi G, Mihaly E, Huang QH, Tatro JB, Rand
WM, Emerson CH, Lechan RM 2000 alpha-Melanocyte-
stimulating hormone is contained in nerve terminals inner-
vating thyrotropin-releasing hormone-synthesizing neurons
in the hypothalamic paraventricular nucleus and prevents
fasting-induced suppression of prothyrotropin-releasing
hormone gene expression. J Neurosci 20:1550–1558.
519. Fekete C, Kelly J, Mihaly E, Sarkar S, Rand WM, Legradi G,
Emerson CH, Lechan RM 2001 Neuropeptide Y has a
central inhibitory action on the hypothalamic-pituitary-
thyroid axis. Endocrinology 142:2606–2613.
520. Fekete C, Sarkar S, Rand WM, Harney JW, Emerson CH,
Bianco AC, Lechan RM 2002 Agouti-related protein
(AGRP) has a central inhibitory action on the hypotha-
lamic-pituitary-thyroid (HPT) axis; comparisons between
the effect of AGRP and neuropeptide Y on energy ho-
meostasis and the HPT axis. Endocrinology 143:3846–3853.
521. Fekete C, Sarkar S, Rand WM, Harney JW, Emerson CH,
Bianco AC, Beck-Sickinger A, Lechan RM 2002 Neuro-
peptide Y1 and Y5 receptors mediate the effects of neuro-
peptide Y on the hypothalamic-pituitary-thyroid axis.
Endocrinology 143:4513–4519.
522. Fekete C, Marks DL, Sarkar S, Emerson CH, Rand WM,
Cone RD, Lechan RM 2004 Effect of Agouti-related protein
in regulation of the hypothalamic-pituitary-thyroid axis in
the melanocortin 4 receptor knockout mouse. Endo-
crinology 145:4816–4821.
523. Garza R, Dussault JH, Puymirat J 1988 Influence of triio-
dothyronine (L-T3) on the morphological and biochemical
development of fetal brain acetylcholinesterase-positive
neurons cultured in a chemically defined medium. Brain
Res 471:287–297.
524. Heuer H, Mason CA 2003 Thyroid hormone induces cere-
bellar Purkinje cell dendritic development via the thyroid
hormone receptor alpha1. J Neurosci 23: 10604–10612.
525. Liu YY, Tachiki KH, Brent GA 2002 A targeted thyroid
hormone receptor alpha gene dominant-negative mutation
(P398H) selectively impairs gene expression in differenti-
ated embryonic stem cells. Endocrinology 143:2664–2672.
526. Mohacsik P, Zeold A, Bianco AC, Gereben B 2011 Thyroid
hormone and the neuroglia: both source and target. J
Thyroid Res 2011:215718.
527. Ruel J, Dussault JH 1985 Triiodothyronine increases glu-
tamine synthetase activity in primary cultures of rat cere-
bellum. Brain Res 353:83–88.
528. Niederkinkhaus V, Marx R, Hoffmann G, Dietzel ID 2009
Thyroid hormone (T3)-induced up-regulation of voltage-
activated sodium current in cultured postnatal hippocam-
pal neurons requires secretion of soluble factors from glial
cells. Mol Endocrinol 23:1494–1504.
529. Emery, B. Regulation of oligodendrocyte differentiation
and myelination. Science 330:779–782.
530. Strait KA, Carlson DJ, Schwartz HL, Oppenheimer JH 1997
Transient stimulation of myelin basic protein gene expres-
sion in differentiating cultured oligodendrocytes: a model
for 3,5,3¢-triiodothyronine-induced brain development.
Endocrinology 138:635–641.
531. Samuels HH, Tsai JS, Cintron R 1973 Thyroid ho rmone
action: a cell-culture system responsive to physiological
concentrations of thyroid hormones. Science 18 1:1253–
1256.
532. Lazar MA 1990 Sodium butyrate selectively alters thyroid
hormone receptor gene expression in GH
3
cells. J Biol
Chem 265:17474–17477.
533. Yusta B, Alarid ET, Gordon DF, Ridgway EC, Mellon PL
1998 The thyrotropin beta-subunit gene is repressed by
thyroid hormone in a novel thyrotrope cell line, mouse T
alphaT1 cells. Endocrinology 139:4476–4482.
534. Mai W, Janier MF, Allioli N, Quignodon L, Chuzel T, Fla-
mant F, Samarut J 2004 Thyroid hormone receptor alpha is
a molecular switch of cardiac function between fetal and
postnatal life. Proc Natl Acad Sci USA 101:10332–10337.
535. Klein I, Ojamaa K 2001 Thyroid hormone and the cardio-
vascular system. N Engl J Med 344:501–509.
536. [Deleted.]
537. Ojamaa K, Samarel AM, Kupfer JM, Hong C, Klein I 1992
Thyroid hormone effects on cardiac gene expression inde-
pendent of cardiac growth and protein synthesis. Am J
Physiol Endocrinol Metab 263:E534–E540.
538. [Deleted.]
539. Thomas TA, Kuzman JA, Anderson BE, Andersen SM,
Schlenker EH, Holder MS, Gerdes AM 2005 Thyroid hor-
mones induce unique and potentially beneficial changes in
cardiac myocyte shape in hypertensive rats near heart
failure. Am J Physiol Heart Circ Physiol 288:H2118–H2122.
540. Henderson KK, Danzi S, Paul JT, Leya G, Klein I, Samarel
AM 2009 Physiological replacement of T
3
improves left
ventricular function in an animal model of myocardial
infarction-induced congestive heart failure. Circ Heart Fail
2:243–252.
541. Suarez J, Scott BT, Suarez-Ramirez JA, Chavira CV, Dill-
mann WH 2010 Thyroid hormone inhibits ERK phos-
phorylation in pressure overload-induced hypertrophied
mouse hearts through a receptor-mediated mechanism. Am
J Physiol Cell Physiol 299:C1524–C1529.
542. van Rooij E, Sutherland LB, Qi X, Richardson JA, Hill J,
Olson EN 2007 Control of stress-dependent cardiac growth
and gene expression by a microRNA. Science 316:575–579.
543. Minakawa M, Takeuchi K, Ito K, Tsushima T, Fukui K,
Takaya S, Fukuda I 2003 Restoration of sarcoplasmic re-
ticulum protein level by thyroid hormone contributes to
partial improvement of myocardial function, but not to
glucose metabolism in an early failing heart. Eur J Cardi-
othorac Surg 24:493–501.
544. Kinugawa K, Yonekura K, Ribeiro RC, Eto Y, Aoyagi T,
Baxter JD, Camacho SA, Bristow MR, Long CS, Simpson
PC 2001 Regulation of thyroid hormone receptor isoforms
in physiological and pathological cardiac hypertrophy. Circ
Res 89:591–598.
545. Hartley CJ, Taffet GE, Reddy AK, Entman ML, Michael LH
2002 Noninvasive cardiovascular phenotyping in mice.
ILAR J 43:147–158.
546. Borst O, Ochmann C, Schonberger T, Jacoby C, Stellos K,
Seizer P, Flogel U, Lang F, Gawaz M 2011 Methods em-
ployed for induction and analysis of experimental myo-
cardial infarction in mice. Cell Physiol Biochem 28:1–12.
160 BIANCO ET AL.
547. Redout EM, van der Toorn A, Zuidwijk MJ, van de Kolk
CW, van Echteld CJ, Musters RJ, van Hardeveld C, Paulus
WJ, Simonides WS 2010 Antioxidant treatment attenuates
pulmonary arterial hypertension-induced heart failure. Am
J Physiol Heart Circ Physiol 298:H1038–H1047.
548. Johansson C, Thoren P 1997 The effects of triiodothyronine
(T3) on heart rate, temperature and ECG measured with
telemetry in freely moving mice. Acta Physiol Scand
160:133–138.
549. Mittag J, Davis B, Vujovic M, Arner A, Vennstrom B 2010
Adaptations of the autonomous nervous system controlling
heart rate are impaired by a mutant thyroid hormone
receptor-alpha1. Endocrinology 151:2388–2395.
550. Van Vliet BN, McGuire J, Chafe L, Leonard A, Joshi A,
Montani JP 2006 Phenotyping the level of blood pressure by
telemetry in mice. Clin Exp Pharmacol Physiol 33:1007–1015.
551. de Waard MC, van der Velden J, Bito V, Ozdemir S, Bies-
mans L, Boontje NM, Dekkers DH, Schoonderwoerd K,
Schuurbiers HC, de Crom R, Stienen GJ, Sipido KR, Lamers
JM, Duncker DJ 2007 Early exercise training normalizes
myofilament function and attenuates left ventricular pump
dysfunction in mice with a large myocardial infarction. Circ
Res 100:1079–1088.
552. Heather LC, Cole MA, Atherton HJ, Coumans WA, Evans
RD, Tyler DJ, Glatz JF, Luiken JJ, Clarke K 2010 Adenosine
monophosphate-activated protein kinase activation, sub-
strate transporter translocation, and metabolism in the con-
tracting hyperthyroid rat heart. Endocrinology 151:422–431.
553. Pantos C, Mourouzis I, Saranteas T, Clave
´
G, Ligeret H,
Noack-Fraissignes P, Renard PY, Massonneau M, Perimenis
P, Spanou D, Kostopanagiotou G, Cokkinos DV 2009 Thyroid
hormone improves postischaemic recovery of function while
limiting apoptosis: a new therapeutic approach to support
hemodynamics in the setting of ischaemia-reperfusion? Basic
Res Cardiol 104:69–77.
554. Beekman RE, van Hardeveld C, Simonides WS 1989 On the
mechanism of the reduction by thyroid hormone of beta-
adrenergic relaxation rate stimulation in rat heart. Biochem
J 259:229–236.
555. Sutherland FJ, Shattock MJ, Baker KE, Hearse DJ 2003
Mouse isolated perfused heart: characteristics and cautions.
Clin Exp Pharmacol Physiol 30:867–878.
556. Hyyti OM, Olson AK, Ge M, Ning XH, Buroker NE, Chung
Y, Jue T, Portman MA 2008 Cardioselective dominant-
negative thyroid hormone receptor (Delta337T) modulates
myocardial metabolism and contractile efficiency. Am J
Physiol Endocrinol Metab 295:E420–E427.
557. Portman MA 2008 Thyroid hormone regulation of heart
metabolism. Thyroid 18:217–225.
558. de Tombe PP, ter Keurs HE 1991 Lack of effect of isopro-
terenol on unloaded velocity of sarcomere shortening in rat
cardiac trabeculae. Circ Res 68:382–391.
559. ter Keurs HE, Deis N, Landesberg A, Nguyen TT, Livshitz
L, Stuyvers B, Zhang ML 2003 Force, sarcomere shortening
velocity and ATPase activity. Adv Exp Med Biol 538:583-
602; discussion 602.
560. Bing OH, Hague NL, Perreault CL, Conrad CH, Brooks
WW, Sen S, Morgan JP 1994 Thyroid hormone effects on
intracellular calcium and inotropic responses of rat ven-
tricular myocardium. Am J Physiol Heart Circ Physiol
267:H1112–H1121.
561. Di Meo S, Venditti P, De Leo T 1997 Effect of iodothyronines
on electrophysiological properties of rat papillary muscle
fibres. Hormone Metabolism Research 29:225–230.
562. Xu Y, Monasky MM, Hiranandani N, Haizlip KM, Billman
GE, Janssen PM 2011 Effect of twitch interval duration on
the contractile function of subsequent twitches in isolated
rat, rabbit, and dog myocardium under physiological
conditions. J Appl Physiol 111: 1159–1167.
563. Szczesna-Cordary D, Jones M, Moore JR, Watt J, Kerrick
WG, Xu Y, Wang Y, Wagg C, Lopaschuk GD 2007 Myosin
regulatory light chain E22K mutation results in decreased
cardiac intracellular calcium and force transients. FASEB J
21:3974–3985.
564. Belke DD, Gloss B, Swanson EA, Dillmann WH 2007
Adeno-associated virus-mediated expression of thyroid
hormone receptor isoforms-alpha1 and -beta1 improves
contractile function in pressure overload-induced cardiac
hypertrophy. Endocrinology 148:2870–2877.
565. Beekman RE, van Hardeveld C, Simonides WS 1990 Thy-
roid status and beta-agonistic effects on cytosolic calcium
concentrations in single rat cardiac myocytes activated by
electrical stimulation or high-K + depolarization. Biochem J
268:563–569.
566. Holt E, Sjaastad I, Lunde PK, Christensen G, Sejersted OM
1999 Thyroid hormone control of contraction and the
Ca(2 + )-ATPase/phospholamban complex in adult rat
ventricular myocytes. J Mol Cell Cardiol 31:645–656.
567. Des Tombe AL, Van Beek-Harmsen BJ, Lee-De Groot MB,
Van Der Laarse WJ 2002 Calibrated histochemistry applied
to oxygen supply and demand in hypertrophied rat myo-
cardium. Microsc Res Tech 58:412–420.
568. Whittaker P, Kloner RA, Boughner DR, Pickering JG 1994
Quantitative assessment of myocardial collagen with pi-
crosirius red staining and circularly polarized light. Basic
Res Cardiol 89:397–410.
569. Pandya K, Kim HS, Smithies O 2006 Fibrosis, not cell size,
delineates beta-myosin heavy chain reexpression during
cardiac hypertrophy and normal aging in vivo. Proc Natl
Acad Sci USA 103:16864–16869.
570. Lopez JE, Myagmar BE, Swigart PM, Montgomery MD,
Haynam S, Bigos M, Rodrigo MC, Simpson PC 2011 beta-
Myosin heavy chain is induced by pressure overload in a
minor subpopulation of smaller mouse cardiac myocytes.
Circ Res 109:629–638.
571. Pantos C, Mourouzis I, Xinaris C, Kokkinos AD, Markakis
K, Dimopoulos A, Panagiotou M, Saranteas T, Kostopa-
nagiotou G, Cokkinos DV 2007 Time-dependent changes in
the expression of thyroid hormone receptor alpha 1 in the
myocardium after acute myocardial infarction: possible
implications in cardiac remodelling. Eur J Endocrinol 156:
415–424.
572. Ojamaa K, Klemperer JD, MacGilvray SS, Klein I, Samarel
A 1996 Thyroid hormone and hemodynamic regulation
of beta-myosin heavy chain promoter in the heart. Endo-
crinology 137:802–808.
573. Sun ZQ, Ojamaa K, Coetzee WA, Artman M, Klein I 2000
Effects of thyroid hormone on action potential and re-
polarizing currents in rat ventricular myocytes. Am J
Physiol Endocrinol Metab 278:E302–E307.
574. Bell D, McDermott BJ 2000 Contribution of de novo protein
synthesis to the hypertrophic effect of IGF-1 but not of
thyroid hormones in adult ventricular cardiomyocytes. Mol
Cell Biochem 206:113–124.
575. Muller A, Zuidwijk MJ, Simonides WS, van Hardeveld C
1997 Modulation of SERCA2 expression by thyroid hor-
mone and norepinephrine in cardiocytes: role of contrac-
tility. Am J Physiol Heart Circ Physiol 272:H1876–H1885.
INVESTIGATING THYROID HORMONE ECONOMY AND ACTION 161
576. Vlasblom R, Muller A, Musters RJ, Zuidwijk MJ, Van
Hardeveld C, Paulus WJ, Simonides WS 2004 Contractile
arrest reveals calcium-dependent stimulation of SERCA2a
mRNA expression in cultured ventricular cardiomyocytes.
Cardiovasc Res 63:537–544.
577. van Dijk-Ottens M, Vos IH, Cornelissen PW, de Bruin A,
Everts ME 2010 Thyroid hormone-induced cardiac me-
chano growth factor expression depends on beating activ-
ity. Endocrinology 151:830–838.
578. Riedel B, Jia Y, Du J, Akerman S, Huang X 2005 Thyroid
hormone inhibits slow skeletal TnI expression in cardiac
TnI-null myocardial cells. Tissue Cell 37:47–51.
579. van der Heide SM, Joosten BJ, Dragt BS, Everts ME, Klaren
PH 2007 A physiological role for glucuronidated thyroid
hormones: preferential uptake by H9c2(2-1) myotubes. Mol
Cell Endocrinol 264:109–117.
580. Pantos C, Xinaris C, Mourouzis I, Perimenis P, Politi E,
Spanou D, Cokkinos DV 2008 Thyroid hormone receptor
alpha 1: a switch to cardiac cell ‘metamorphosis’’? J Physiol
Pharmacol 59:253–269.
581. Meischl C, Buermans HP, Hazes T, Zuidwijk MJ, Musters
RJ, Boer C, van Lingen A, Simonides WS, Blankenstein MA,
Dupuy C, Paulus WJ, Hack CE, Ris-Stalpers C, Roos D,
Niessen HW 2008 H9c2 cardiomyoblasts produce thyroid
hormone. Am J Physiol Cell Physiol 294:C1227–C1233.
582. McAllister RM, Grossenburg VD, Delp MD, Laughlin MH
1998 Effects of hyperthyroidism on vascular contractile
and relaxation responses. Am J Physiol Endocrinol Metab
274:E946–E953.
583. Virdis A, Colucci R, Fornai M, Polini A, Daghini E, Duranti
E, Ghisu N, Versari D, Dardano A, Blandizzi C, Taddei S, Del
Tacca M, Monzani F 2009 Inducible nitric oxide synthase is
involved in endothelial dysfunction of mesenteric small ar-
teries from hypothyroid rats. Endocrinology 150:1033–1042.
584. Deng J, Zhao R, Zhang Z, Wang J 2010 Changes in vasor-
eactivity of rat large- and medium-sized arteries induced
by hyperthyroidism. Exp Toxicol Pathol 62:317–322.
585. Pappas M, Mourouzis K, Karageorgiou H, Tesseromatis C,
Mourouzis I, Kostopanagiotou G, Pantos C, Cokkinos DV
2009 Thyroid hormone modulates the responsiveness of rat
aorta to alpha1-adrenergic stimulation: an effect due to
increased activation of beta2-adrenergic signaling. Int An-
giol 28:474–478.
586. Ojamaa K, Klemperer JD, Klein I 1996 Acute effects of thyroid
hormone on vascular smooth muscle. Thyroid 6:505–512.
587. Carrillo-Sepulveda MA, Ceravolo GS, Fortes ZB, Carvalho
MH, Tostes RC, Laurindo FR, Webb RC, Barreto-Chaves
ML 2010 Thyroid hormone stimulates NO production via
activation of the PI3K/Akt pathway in vascular myocytes.
Cardiovasc Res 85:560–570.
588. Bianco AC, Maia AL, da Silva WS, Christoffolete MA 2005
Adaptive activation of thyroid hormone and energy ex-
penditure. Biosci Rep 25:191–208.
589. Silva JE 2006 Thermogenic mechanisms and their hormonal
regulation. Physiol Rev 86:435–464.
590. Sestoft L 1980 Metabolic aspects of the calorigenic effect of
thyroid hormone in mammals. Clin Endocrinol 13:489–506.
591. [Deleted.]
592. Rabelo R, Schifman A, Rubio A, Sheng X, Silva JE 1995
Delineation of thyroid hormone-responsive sequences
within a critical enhancer in the rat uncoupling protein
gene. Endocrinology 136:1003–1013.
593. Curcio C, Lopes AM, Ribeiro MO, Francoso OA Jr, Car-
valho SD, Lima FB, Bicudo JE, Bianco AC 1999 Develop-
ment of compensatory thermogenesis in response to
overfeeding in hypothyroid rats. Endocrinology 140:3438–
3443.
594. Hollenberg AN 2008 The role of the thyrotropin-releasing
hormone (TRH) neuron as a metabolic sensor. Thyroid
18:131–139.
595. Lechan RM, Fekete C 2006 The TRH neuron: a hypotha-
lamic integrator of energy metabolism. Prog Brain Res
153:209–235.
596. Freake HC, Schwartz HL, Oppenheimer JH 1989 The reg-
ulation of lipogenesis by thyroid hormone and its contri-
bution to thermogenesis. Endocrinology 125:2868–2874.
597. Myant NB, Witney S 1967 The time course of the effect of
thyroid hormones upon basal oxygen consumption and
plasma concentration of free fatty acid in rats. J Physiol
190:221–228.
598. Himsworth RL 1960 Effect of adrenaline and insulin upon
the oxygen consumption of hyperthyroid rats. Nature
185:694.
599. Smith DC, Brown FC 1952 The effect of parrot fish thyroid
extract on the respiratory metabolism of the white rat. Biol
Bull 102:278–286.
600. Oppenheimer JH, Schwartz HL, Lane JT, Thompson MP
1991 Functional relationship of thyroid hormone-induced
lipogenesis, lipolysis, and thermogenesis. J Clin Invest
87:125–132.
601. Taylor CR, Schmidt-Nielsen K, Raab JL 1970 Scaling of
energetic cost of running to body size in mammals. Am J
Physiol 219:1104–1107.
602. Christoffolete MA, Linardi CC, de Jesus L, Ebina KN,
Carvalho SD, Ribeiro MO, Rabelo R, Curcio C, Martins L,
Kimura ET, Bianco AC 2004 Mice with targeted disruption
of the Dio2 gene have cold-induced overexpression of the
uncoupling protein 1 gene but fail to increase brown adi-
pose tissue lipogenesis and adaptive thermogenesis. Dia-
betes 53:577–584.
603. Villicev CM, Freitas FR, Aoki MS, Taffarel C, Scanlan TS,
Moriscot AS, Ribeiro MO, Bianco AC, Gouveia CH 2007
Thyroid hormone receptor beta-specific agonist GC-1 in-
creases energy expenditure and prevents fat-mass accu-
mulation in rats. J Endocrinol 193:21–29.
604. MacLagan NF, Sheahan MM 1950 The measurement of
oxygen consumption in small animals by a closed circuit
method. J Endocrinol 6:456–462.
605. [Deleted.]
606. Miller CN, Kauffman TG, Cooney PT, Ramseur KR, Brown
LM 2010 Comparison of DEXA and QMR for assessing
fat and lean body mass in adult rats. Physiol Behav 103:
117–121.
607. Tscho
¨
p MH, Speakman JR, Arch JR, Auwerx J, Bru
¨
ning JC,
Chan L, Eckel RH, Farese RV Jr, Galgani JE, Hambly C,
Herman MA, Horvath TL, Kahn BB, Kozma SC, Maratos-
Flier E, Mu
¨
ller TD, Mu
¨
nzberg H, Pfluger PT, Plum L, Re-
itman ML, Rahmouni K, Shulman GI, Thomas G, Kahn CR,
Ravussin E 2011 A guide to analysis of mouse energy
metabolism. Nat Methods 9:57–63.
608. Sjo
¨
gren M, Alkemade A, Mittag J, Nordstro
¨
m K, Katz A,
Rozell B, Westerblad H, Arner A, Vennstro
¨
m B 2007 Hy-
permetabolism in mice caused by the central action of an
unliganded thyroid hormone receptor alpha1. EMBO J
26:4535–4545.
609. Yang Y, Gordon CJ 1997 Regulated hypothermia in the hy-
pothyroid rat induced by administration of propylthiouracil.
Am J Physiol Regul Integr Comp Physiol 272:R1390–R1395.
162 BIANCO ET AL.
610. Gordon CJ, Becker P, Padnos B 2000 Comparison of heat
and cold stress to assess thermoregulatory dysfunction in
hypothyroid rats. Am J Physiol Regul Integr Comp Physiol
279:R2066–R2071.
611. Kasson BG, George R 1983 Thermoregulation in hyper-
thyroid rats: mechanism underlying the lack of hypother-
mic response to morphine in hyperthyroid animals. Life Sci
33:1845–1852.
612. Harper ME, Ballantyne JS, Leach M, Brand MD 1993 Effects
of thyroid hormones on oxidative phosphorylation. Bio-
chem Soc Trans 21(Pt 3):785–792.
613. Harper ME, Brand MD 1995 Use of top-down elasticity
analysis to identify sites of thyroid hormone-induced
thermogenesis. Proc Soc Exp Biol Med 208:228–237.
614. Harper ME, Brand MD 1993 The quantitative contribu-
tions of mitochondrial proton leak and ATP turnover re-
actions to the changed respiration rates of hepatocytes from
rats of different thyroid status. J Biol Chem 268:14850–
14860.
615. Ferrick DA, Neilson A, Beeson C 2008 Advances in mea-
suring cellular bioenergetics using extracellular flux. Drug
Discov Today 13:268–274.
616. Wu M, Neilson A, Swift AL, Moran R, Tamagnine J, Par-
slow D, Armistead S, Lemire K, Orrell J, Teich J, Chomicz S,
Ferrick DA 2007 Multiparameter metabolic analysis reveals
a close link between attenuated mitochondrial bioenergetic
function and enhanced glycolysis dependency in human
tumor cells. Am J Physiol Cell Physiol 292:C125–C136.
617. Yuan C, Lin JZ, Sieglaff DH, Ayers SD, Denoto-Reynolds F,
Baxter JD, Webb P 2012 Identical gene regulation patterns
of T
3
and selective thyroid hormone receptor modulator
GC-1. Endocrinology 153:501–511.
618. Sukocheva OA, Carpenter DO 2006 Anti-apoptotic effects
of 3,5,3¢-tri-iodothyronine in mouse hepatocytes. J En-
docrinol 191:447–458.
619. Giudetti AM, Leo M, Geelen MJ, Gnoni GV 2005 Short-term
stimulation of lipogenesis by 3,5-L-diiodothyronine in cul-
tured rat hepatocytes. Endocrinology 146:3959–3966.
620. Hafner RP, Brown GC, Brand MD 1990 Thyroid-hormone
control of state-3 respiration in isolated rat liver mito-
chondria. Biochem J 265:731–734.
621. Hafner RP, Nobes CD, McGown AD, Brand MD 1988 Al-
tered relationship between protonmotive force and respi-
ration rate in non-phosphorylating liver mitochondria
isolated from rats of different thyroid hormone status. Eur J
Biochem 178:511–518.
622. Carvalho SD, Bianco AC, Silva JE 1996 Effects of hypo-
thyroidism on brown adipose tissue adenylyl cyclase ac-
tivity. Endocrinology 137:5519–5529.
623. Pachucki J, Hopkins J, Peeters R, Tu H, Carvalho SD,
Kaulbach H, Abel ED, Wondisford FE, Ingwall JS, Larsen
PR 2001 Type 2 iodothyronine deiodinase transgene ex-
pression in the mouse heart causes cardiac-specific thyro-
toxicosis. Endocrinology 142:13–20.
624. Kim B, Carvalho-Bianco SD, Larsen PR 2004 Thyroid hor-
mone and adrenergic signaling in the heart. Arq Bras En-
docrinol Metabol 48:171–175.
625. Ring GC 1942 The importance of the thyroid in maintaining
an adequate production of heat during exposure to cold.
Am J Physiol 137:582–588.
626. Sellers EA, You SS 1950 Role of the thyroid in metabolic
responses to a cold environment. Am J Physiol 163:81–91.
627. Ribeiro MO, Lebrun FL, Christoffolete MA, Branco M,
Crescenzi A, Carvalho SD, Negrao N, Bianco AC 2000
Evidence of UCP1-independent regulation of norepineph-
rine-induced thermogenesis in brown fat. Am J Physiol
Endocrinol Metab 279:E314–E322.
628. [Deleted.]
629. Branco M, Ribeiro M, Negrao N, Bianco AC 1999 3,5,3¢-
Triiodothyronine actively stimulates UCP in brown fat
under minimal sympathetic activity. Am J Physiol Endo-
crinol Metab 276:E179–E187.
630. Silva JE 2000 Catecholamines and the sympathoadrenal
system in hypothyroidism. In: Braverman LE, Utiger RD
(eds) Werner & Ingbar’s The Thyroid. A Fundamental and
Clinical Text, 8th edition. Lippincott Williams & Wilkins,
Philadelphia, pp 820–823.
631. Silva JE 2000 Catecholamines and the sympathoadrenal
system in thyrotoxicosis. In: Braverman LE, Utiger RD (eds)
Werner & Ingbar’s The Thyroid. A Fundamental and
Clinical Text, 8th edition. Lippincott Williams & Wilkins,
Philadelphia, pp 642–651.
632. Landsberg L 1977 Catecholamines and hyperthyroidism.
Clin Endocrinol Metab 6:697–718.
633. Young JB, Saville E, Landsberg L 1982 Effect of thyroid
state on norepinephrine (NE) turnover in rat brown adi-
pose tissue (BAT): potential importance of the pituitary.
Clin Res 32:407 (Abstract).
634. Rothwell NJ, Stock MJ 1979 A role for brown adipose tissue
in diet-induced thermogenesis. Nature 281:31–35.
635. Syed MA, Thompson MP, Pachucki J, Burmeister LA 1999
The effect of thyroid hormone on size of fat depots accounts
for most of the changes in leptin mRNA and serum levels
in the rat. Thyroid 9:503–512.
636. [Deleted.]
637. Freake HC, Moon YK 2003 Hormonal and nutritional reg-
ulation of lipogenic enzyme mRNA levels in rat primary
white and brown adipocytes. J Nutr Sci Vitaminol (Tokyo)
49:40–46.
638. Silva JE 1988 Full expression of uncoupling protein gene
requires the concurrence of norepinephrine and triiodo-
thyronine. Mol Endocrinol 2:706–713.
639. [Deleted.]
640. Carvalho SD, Negrao N, Bianco AC 1993 Hormonal regulation
of malic enzyme and glucose-6-phosphate dehydrogenase
in brown adipose tissue. Am J Physiol Endocrinol Metab
264:E874–E881.
641. Antinozzi PA, Segall L, Prentki M, McGarry JD, Newgard
CB 1998 Molecular or pharmacologic perturbation of the
link between glucose and lipid metabolism is without effect
on glucose-stimulated insulin secretion. A re-evaluation of
the long-chain acyl-CoA hypothesis. J Biol Chem 273:
16146–16154.
642. Saggerson ED, McAllister TW, Baht HS 1988 Lipogenesis in
rat brown adipocytes. Effects of insulin and noradrenaline,
contributions from glucose and lactate as precursors and
comparisons with white adipocytes. Biochem J 251:701–709.
643. Bachman ES, Dhillon H, Zhang CY, Cinti S, Bianco AC,
Kobilka BK, Lowell BB 2002 betaAR signaling required for
diet-induced thermogenesis and obesity resistance. Science
297:843–845.
644. Wang YX, Lee CH, Tiep S, Yu RT, Ham J, Kang H, Evans
RM 2003 Peroxisome-proliferator-activated receptor delta
activates fat metabolism to prevent obesity. Cell 113:159–
170.
645. Hellstro
¨
m L, Wahrenberg H, Reynisdottir S, Arner P 1997
Catecholamine-induced adipocyte lipolysis in human hy-
perthyroidism. J Clin Endocrinol Metab 82:159–166.
INVESTIGATING THYROID HORMONE ECONOMY AND ACTION 163
646. Lonnqvist F, Wahrenberg H, Hellstrom L, Reynisdottir S,
Arner P 1992 Lipolytic catecholamine resistance due to
decreased beta 2-adrenoceptor expression in fat cells. J Clin
Invest 90:2175–2186.
647. Wahrenberg H, Lonnqvist F, Arner P 1989 Mechanisms
underlying regional differences in lipolysis in human adi-
pose tissue. J Clin Invest 84:458–467.
648. Bianco AC, Kieffer JD, Silva JE 1992 Adenosine 3¢,5¢-
monophosphate and thyroid hormone control of uncoupling
protein messenger ribonucleic acid in freshly dispersed
brown adipocytes. Endocrinology 130:2625–2633.
649. Rehnmark S, Bianco AC, Kieffer JD, Silva JE 1992 Tran-
scriptional and posttranscriptional mechanisms in un-
coupling protein mRNA response to cold. Am J Physiol
Endocrinol Metab 262:E58–E67.
650. [Deleted.]
651. Ying H, Araki O, Furuya F, Kato Y, Cheng SY 2007 Im-
paired adipogenesis caused by a mutated thyroid hormone
alpha1 receptor. Mol Cell Biol 27:2359–2371.
652. Obregon MJ 2008 Thyroid hormone and adipocyte differ-
entiation. Thyroid 18:185–195.
653. Benito M, Porras A, Santos E 1993 Establishment of per-
manent brown adipocyte cell lines achieved by transfection
with SV40 large T antigen and ras genes. Exp Cell Res
209:248–254.
654. Surks MI, Oppenheimer JH 1977 Concentration of L-
thyroxine and L-triiodothyronine specifically bound to
nuclear receptors in rat liver and kidney. Quantitative
evidence favoring a major role of T
3
in thyroid hormone
action. J Clin Invest 60:555–562.
655. Carr FE, Jump DB, Oppenheimer JH 1984 Distribution of
thyroid hormone-responsive translated products in rat liver
polysome and postribosomal ribonucleoprotein popula-
tions. Endocrinology 115:1737–1745.
656. Ruegamer WR, Newman GH, Richert DA, Westerfeld WW
1965 Specificity of the alpha-glycerophosphate dehydro-
genase and malic enzyme response to thyroxine. En-
docrinology 77:707–715.
657. Mariash CN, Kaiser FE, Schwartz HL, Towle HC, Oppen-
heimer JH 1980 Synergism of thyroid hormone and high
carbohydrate diet in the induction of lipogenic enzymes in the
rat. Mechanisms and implications. J Clin Invest 65:1126–1134.
658. Towle HC, Mariash CN, Oppenheimer JH 1980 Changes in
the hepatic levels of messenger ribonucleic acid for malic
enzyme during induction by thyroid hormone or diet.
Biochemistry 19:579–585.
659. Mariash CN, Kaiser FE, Oppenheimer JH 1980 Comparison
of the response characteristics of four lipogenic enzymes to
3,5,3¢-triiodothyronine administration: evidence for vari-
able degrees of amplification of the nuclear 3,5,3¢-triiodo-
thyronine signal. Endocrinology 106:22–27.
660. Stakkestad JA, Bremer J 1983 The outer carnitine palmi-
toyltransferase and regulation of fatty acid metabolism in
rat liver in different thyroid states. Biochim Biophys Acta
750:244–252.
661. Stakkestad JA, Bremer J 1982 The metabolism of fatty acids
in hepatocytes isolated from triiodothyronine-treated rats.
Biochim Biophys Acta 711:90–100.
662. Heimberg M, Olubadewo JO, Wilcox HG 1985 Plasma li-
poproteins and regulation of hepatic metabolism of fatty
acids in altered thyroid states. Endocr Rev 6:590–607.
663. Engelken SF, Eaton RP 1980 Thyroid hormone-induced
dissociation between plasma triglyceride and cholesterol
regulation in the rat. Endocrinology 107:208–214.
664. Keyes WG, Heimberg M 1979 Influence of thyroid status on
lipid metabolism in the perfused rat liver. J Clin Invest
64:182–190.
665. Apostolopoulos JJ, Howlett GJ, Fidge N 1987 Effects of
dietary cholesterol and hypothyroidism on rat apolipo-
protein mRNA metabolism. J Lipid Res 28:642–648.
666. Dolphin PJ, Forsyth SJ 1983 Nascent hepatic lipoproteins in
hypothyroid rats. J Lipid Res 24:541–551.
667. Dolphin PJ 1981 Serum and hepatic nascent lipoproteins
in normal and hypercholesterolemic rats. J Lipid Res 22:
971–989.
668. Johansson L, Rudling M, Scanlan TS, Lundasen T, Webb P,
Baxter J, Angelin B, Parini P 2005 Selective thyroid receptor
modulation by GC-1 reduces serum lipids and stimulates
steps of reverse cholesterol transport in euthyroid mice.
Proc Natl Acad Sci USA 102:10297–10302.
669. Ribeiro MO 2008 Effects of thyroid hormone analogs on
lipid metabolism and thermogenesis. Thyroid 18:197–203.
670. Amorim BS, Ueta CB, Freitas BC, Nassif RJ, Gouveia CH,
Christoffolete MA, Moriscot AS, Lancelloti CL, Llimona F,
Barbeiro HV, de Souza HP, Catanozi S, Passarelli M, Aoki
MS, Bianco AC, Ribeiro MO 2009 A TRbeta-selective ago-
nist confers resistance to diet-induced obesity. J Endocrinol
203:291–299.
671. Simonides WS, van Hardeveld C 2008 Thyroid hormone as
a determinant of metabolic and contractile phenotype of
skeletal muscle. Thyroid 18:205–216.
672. Simonides WS, Thelen MH, van der Linden CG, Muller A,
van Hardeveld C 2001 Mechanism of thyroid-hormone
regulated expression of the SERCA genes in skeletal mus-
cle: implications for thermogenesis. Biosci Rep 21:139–154.
673. Jackman MR, Willis WT 1996 Characteristics of mitochon-
dria isolated from type I and type IIb skeletal muscle. Am J
Physiol Cell Physiol 270:C673–C678.
674. Pette D, Staron RS 1997 Mammalian skeletal muscle fiber
type transitions. Int Rev Cytol 170:143–223.
675. Simonides WS, van Hardeveld C 1989 The postnatal de-
velpment of sarcoplasmic reticulum Ca
2 +
-transport activity
in skeletal muscle of the rat is critically dependent on
thyroid hormone. Endocrinology 124:1145–1153.
676. van der Linden CG, Simonides WS, Muller A, van der
Laarse WJ, Vermeulen JL, Zuidwijk MJ, Moorman AF, van
Hardeveld C 1996 Fiber-specific regulation of Ca(2 + )-AT-
Pase isoform expression by thyroid hormone in rat skeletal
muscle. Am J Physiol Cell Physiol 271:C1908–C1919.
677. Butler-Browne GS, Herlicoviez D, Whalen RG 1984 Effects
of hypothyroidism on myosin isozyme transitions in de-
veloping rat muscle. FEBS Lett 166:71–75.
678. Mahdavi V, Izumo S, Nadal-Ginard B 1987 Developmental
and hormonal regulation of sarcomeric myosin heavy chain
gene family. Circ Res 60:804–814.
679. d’Albis A, Chanoine C, Janmot C, Mira JC, Couteaux R
1990 Muscle-specific response to thyroid hormone of my-
osin isoform transitions during rat postnatal development.
Eur J Biochem 193:155–161.
680. Izumo S, Nadal-Ginard B, Mahdavi V 1986 All members of
the MHC multigene family respond to thyroid hormone in
a highly tissue-specific manner. Science 231:597–600.
681. Ariano MA, Armstrong RB, Edgerton VR 1973 Hindlimb
muscle fiber populations of five mammals. J Histochem
Cytochem 21:51–55.
682. Wang LC, Kernell D 2001 Fibre type regionalisation in
lower hindlimb muscles of rabbit, rat and mouse: a com-
parative study. J Anat 199:631–643.
164 BIANCO ET AL.
683. Baker DJ, Hepple RT 2005 The versatility of the pump-
perfused rat hindlimb preparation: examples relating to
skeletal muscle function and energy metabolism. Can J
Appl Physiol 30:576–590.
684. McAllister RM, Ogilvie RW, Terjung RL 1991 Functional
and metabolic consequences of skeletal muscle remodeling
in hypothyroidism. Am J Physiol Endocrinol Metab 260:
E272–E279.
685. Everts ME, van Hardeveld C, Ter Keurs HE, Kassenaar AA
1983 Force development and metabolism in perfused
skeletal muscle of euthyroid and hyperthyroid rats. Horm
Metab Res 15:388–393.
686. Everts ME, van Hardeveld C, Ter Keurs HE, Kassenaar AA
1981 Force development and metabolism in skeletal muscle
of euthyroid and hypothyroid rats. Acta Endocrinol (Co-
penh) 97:221–225.
687. Kraus B, Pette D 1997 Quantification of MyoD, myogenin,
MRF4 and Id-1 by reverse-transcriptase polymerase chain
reaction in rat muscles–effects of hypothyroidism and chronic
low-frequency stimulation. Eur J Biochem 247:98–106.
688. Kirschbaum BJ, Kucher HB, Termin A, Kelly AM, Pette D
1990 Antagonistic effects of chronic low frequency stimu-
lation and thyroid hormone on myosin expression in rat
fast-twitch muscle. J Biol Chem 265:13974–13980.
689. Montgomery A 1992 The time course of thyroid-hormone-
induced changes in the isotonic and isometric properties of
rat soleus muscle. Pflugers Arch 421:350–356.
690. Everts ME, Simonides WS, Leijendekker WJ, van Hard-
eveld C 1987 Fatigability and recovery of rat soleus muscle
in hyperthyroidism. Metabolism 36:444–450.
691. Johansson C, Lannergren J, Lunde PK, Vennstrom B, Thoren
P, Westerblad H 2000 Isometric force and endurance in so-
leus muscle of thyroid hormone receptor-alpha(1)- or -beta-
deficient mice. Am J Physiol Regul Integr Comp Physiol
278:R598–R603.
692. Li X, Hughes SM, Salviati G, Teresi A, Larsson L 1996
Thyroid hormone effects on contractility and myosin
composition of soleus muscle and single fibres from young
and old rats. J Physiol 494(Pt 2):555–567.
693. Fitts RH, Brimmer CJ, Troup JP, Unsworth BR 1984 Con-
tractile and fatigue properties of thyrotoxic rat skeletal
muscle. Muscle Nerve 7:470–477.
694. Ottenheijm CA, Hidalgo C, Rost K, Gotthardt M, Granzier
H 2009 Altered contractility of skeletal muscle in mice de-
ficient in titin’s M-band region. J Mol Biol 393:10–26.
695. Norenberg KM, Herb RA, Dodd SL, Powers SK 1996 The
effects of hypothyroidism on single fibers of the rat soleus
muscle. Can J Physiol Pharmacol 74:362–367.
696. van Hees HW, Ottenheijm CA, Granzier HL, Dekhuijzen PN,
Heunks LM 2010 Heart failure decreases passive tension
generation of rat diaphragm fibers. Int J Cardiol 141:275–283.
697. Wahrmann JP, Fulla Y, Rieu M, Kahn A, Dinh-Xuan AT
2002 Altered myosin isoform expression in rat skeletal
muscles induced by a changed thyroid state. Acta Physiol
Scand 176:233–243.
698. Yamada T, Mishima T, Sakamoto M, Sugiyama M, Mat-
sunaga S, Wada M 2007 Myofibrillar protein oxidation and
contractile dysfunction in hyperthyroid rat diaphragm. J
Appl Physiol 102:1850–1855.
699. Agbulut O, Noirez P, Beaumont F, Butler-Browne G 2003
Myosin heavy chain isoforms in postnatal muscle devel-
opment of mice. Biol Cell 95:399–406.
700. Miyabara EH, Aoki MS, Soares AG, Saltao RM, Vilicev CM,
Passarelli M, Scanlan TS, Gouveia CH, Moriscot AS 2005
Thyroid hormone receptor-beta-selective agonist GC-24
spares skeletal muscle type I to II fiber shift. Cell Tissue Res
321:233–241.
701. Yamada T, Inashima S, Matsunaga S, Nara I, Kajihara H,
Wada M 2004 Different time course of changes in sarco-
plasmic reticulum and myosin isoforms in rat soleus mus-
cle at early stage of hyperthyroidism. Acta Physiol Scand
180:79–87.
702. Yu F, Gothe S, Wikstrom L, Forrest D, Vennstrom B,
Larsson L 2000 Effects of thyroid hormone receptor gene
disruption on myosin isoform expression in mouse skeletal
muscles. Am J Physiol Regul Integr Comp Physiol 278:
R1545–R1554.
703. dos Santos RA, Giannocco G, Nunes MT 2001 Thyroid
hormone stimulates myoglobin expression in soleus and
extensorum digitalis longus muscles of rats: concomitant
alterations in the activities of Krebs cycle oxidative en-
zymes. Thyroid 11:545–550.
704. Irrcher I, Adhihetty PJ, Sheehan T, Joseph AM, Hood DA
2003 PPARgamma coactivator-1alpha expression during
thyroid hormone- and contractile activity-induced mito-
chondrial adaptations. Am J Physiol Cell Physiol 284:
C1669–C1677.
705. Simonides WS, van Hardeveld C 1990 An assay for sarco-
plasmic reticulum Ca2( + )-ATPase activity in muscle ho-
mogenates. Anal Biochem 191:321–331.
706. Silvestri E, Moreno M, Lombardi A, Ragni M, de Lange P,
Alexson SE, Lanni A, Goglia F 2005 Thyroid-hormone ef-
fects on putative biochemical pathways involved in UCP3
activation in rat skeletal muscle mitochondria. FEBS Lett
579:1639–1645.
707. Venditti P, Puca A, Di Meo S 2003 Effect of thyroid state on
rate and sites of H2O
2
production in rat skeletal muscle
mitochondria. Arch Biochem Biophys 411:121–128.
708. Schiaffino S, Gorza L, Sartore S, Saggin L, Ausoni S, Via-
nello M, Gundersen K, Lomo T 1989 Three myosin heavy
chain isoforms in type 2 skeletal muscle fibres. J Muscle Res
Cell Motil 10:197–205.
709. Smerdu V, Soukup T 2008 Demonstration of myosin heavy
chain isoforms in rat and humans: the specificity of seven
available monoclonal antibodies used in immunohisto-
chemical and immunoblotting methods. Eur J Histochem
52:179–190.
710. Bloemberg D, Quadrilatero J 2012 Rapid determination of
myosin heavy chain expression in rat, mouse, and human
skeletal muscle using multicolor immunofluorescence
analysis. PLoS One 7:e35273.
711. Nunes MT, Bianco AC, Migala A, Agostini B, Hasselbach
W 1985 Thyroxine induced transformation in sarcoplasmic
reticulum of rabbit soleus and psoas muscles. Z Nat-
urforsch [C] 40:726–734.
712. Naumann K, Pette D 1994 Effects of chronic stimulation with
different impulse patterns on the expression of myosin iso-
forms in rat myotube cultures. Differentiation 55:203–211.
713. Grozovsky R, Ribich S, Rosene ML, Mulcahey MA, Huang
SA, Patti ME, Bianco AC, Kim BW 2009 Type 2 deiodinase
expression is induced by peroxisomal proliferator-activated
receptor-gamma agonists in skeletal myocytes. Endocrino-
logy 150:1976–1983.
714. Rando TA, Blau HM 1994 Primary mouse myoblast puri-
fication, characterization, and transplantation for cell-
mediated gene therapy. J Cell Biol 125:1275–1287.
715. Muller A, Thelen MH, Zuidwijk MJ, Simonides WS, van
Hardeveld C 1996 Expression of MyoD in cultured primary
INVESTIGATING THYROID HORMONE ECONOMY AND ACTION 165
myotubes is dependent on contractile activity: correlation
with phenotype-specific expression of a sarcoplasmic re-
ticulum Ca(2 + )-ATPase isoform. Biochem Biophys Res
Commun 229:198–204.
716. Marsili A, Tang D, Harney JW, Singh P, Zavacki AM,
Dentice M, Salvatore D, Larsen PR 2011 Type II iodothyr-
onine deiodinase provides intracellular 3,5,3¢-triiodothyro-
nine to normal and regenerating mouse skeletal muscle.
Am J Physiol Endocrinol Metab 301:E818–E824.
717. Sultan KR, Henkel B, Terlou M, Haagsman HP 2006
Quantification of hormone-induced atrophy of large myo-
tubes from C2C12 and L6 cells: atrophy-inducible and at-
rophy-resistant C2C12 myotubes. Am J Physiol Cell Physiol
290:C650–C659.
718. Thelen MH, Simonides WS, van Hardeveld C 1997 Elec-
trical stimulation of C2C12 myotubes induces contractions
and represses thyroid-hormone-dependent transcription of
the fast-type sarcoplasmic-reticulum Ca2 + -ATPase gene.
Biochem J 321(Pt 3):845–848.
719. Nagase I, Yoshida S, Canas X, Irie Y, Kimura K, Yoshida T,
Saito M 1999 Up-regulation of uncoupling protein 3 by
thyroid hormone, peroxisome proliferator-activated recep-
tor ligands and 9-cis retinoic acid in L6 myotubes. FEBS
Lett 461:319–322.
720. Thelen MH, Simonides WS, Muller A, van Hardeveld C
1998 Cross-talk between transcriptional regulation by
thyroid hormone and myogenin: new aspects of the Ca2 + -
dependent expression of the fast-type sarcoplasmic reticu-
lum Ca2 + -ATPase. Biochem J 329(Pt 1):131–136.
721. Muller A, van Hardeveld C, Simonides WS, van Rijn J 1991
The elevation of sarcoplasmic reticulum Ca2( + )-ATPase
levels by thyroid hormone in the L6 muscle cell line is
potentiated by insulin-like growth factor-1. Biochem J
275:35–40.
722. [Deleted.]
723. Connor MK, Irrcher I, Hood DA 2001 Contractile activity-
induced transcriptional activation of cytochrome C in-
volves Sp1 and is proportional to mitochondrial ATP syn-
thesis in C2C12 muscle cells. J Biol Chem 276:15898–15904.
724. Bassett JH, Williams GR 2008 Critical role of the hypotha-
lamic-pituitary-thyroid axis in bone. Bone 43:418–426.
725. Bassett JH, Boyde A, Howell PG, Bassett RH, Galliford TM,
Archanco M, Evans H, Lawson MA, Croucher P, St Germain
DL, Galton VA, Williams GR 2010 Optimal bone strength
and mineralization requires the type 2 iodothyronine deio-
dinase in osteoblasts. Proc Natl Acad Sci USA 107:7604–7609.
726. Bassett JH, Nordstrom K, Boyde A, Howell PG, Kelly S,
Vennstrom B, Williams GR 2007 Thyroid status during
skeletal development determines adult bone structure and
mineralization. Mol Endocrinol 21:1893–1904.
727. Bassett JH, O’Shea PJ, Sriskantharajah S, Rabier B, Boyde A,
HowellPG,WeissRE,RouxJP,Malaval L, Clement-Lacroix
P, Samarut J, Chassande O, Williams GR 2007 Thyroid hor-
mone excess rather than thyrotropin deficiency induces os-
teoporosis in hyperthyroidism. Mol Endocrinol 21:1095–1107.
728. Bassett JH, van der Spek A, Gogakos A, Williams GR 2012
Quantitative X-ray imaging of rodent bone by Faxitron.
Methods Mol Biol 816:499–506.
729. Bassett JH, Williams AJ, Murphy E, Boyde A, Howell PG,
Swinhoe R, Archanco M, Flamant F, Samarut J, Costagliola
S, Vassart G, Weiss RE, Refetoff S, Williams GR 2008 A lack
of thyroid hormones rather than excess thyrotropin causes
abnormal skeletal development in hypothyroidism. Mol
Endocrinol 22:501–512.
730. O’Shea PJ, Bassett JH, Sriskantharajah S, Ying H, Cheng SY,
Williams GR 2005 Contrasting skeletal phenotypes in mice
with an identical mutation targeted to thyroid hormone
receptor alpha1 or beta. Mol Endocrinol 19:3045–3059.
731. O’Shea PJ, Harvey CB, Suzuki H, Kaneshige M, Kaneshige
K, Cheng SY, Williams GR 2003 A thyrotoxic skeletal
phenotype of advanced bone formation in mice with re-
sistance to thyroid hormone. Mol Endocrinol 17:1410–1424.
732. Esapa CT, Hough TA, Testori S, Head RA, Crane EA, Chan
CP, Evans H, Bassett JH, Tylzanowski P, McNally EG,
Carr AJ, Boyde A, Howell PG, Clark A, Williams GR,
Brown MA, Croucher PI, Nesbit MA, Brown SD, Cox RD,
Cheeseman MT, Thakker RV 2012 A mouse model for
spondyloepiphyseal dysplasia congenita with secondary
osteoarthritis due to a Col2a1 mutation. J Bone Miner Res
27:413–428.
733. Karunaratne A, Esapa C, Hiller J, Boyde A, Head R, Bassett
J, Terrill N, Williams G, Brown M, Croucher P, Brown SD,
Cox RD, Barber AH, Thakker RV, Gupta HS 2012 Sig-
nificant deterioration in nanomechanical quality occurs
through incomplete extrafibrillar mineralization in rachitic
bone: evidence from in-situ synchrotron X-ray scattering
and backscattered electron imaging. J Bone Miner Res
27:876–890.
734. Vanleene M, Saldanha Z, Cloyd KL, Jell G, Bou-Gharios G,
Bassett JH, Williams GR, Fisk NM, Oyen ML, Stevens MM,
Guillot PV, Shefelbine SJ 2011 Transplantation of human
fetal blood stem cells in the osteogenesis imperfecta mouse
leads to improvement in multiscale tissue properties. Blood
117:1053–1060.
735. Peltoketo H, Strauss L, Karjalainen R, Zhang M, Stamp
GW, Segaloff DL, Poutanen M, Huhtaniemi IT 2010 Female
mice expressing constitutively active mutants of FSH re-
ceptor present with a phenotype of premature follicle de-
pletion and estrogen excess. Endocrinology 151:1872–1883.
736. Parfitt AM, Drezner MK, Glorieux FH, Kanis JA, Malluche
H, Meunier PJ, Ott SM, Recker RR 1987 Bone histomor-
phometry: standardization of nomenclature, symbols, and
units. Report of the ASBMR Histomorphometry Nomen-
clature Committee. J Bone Miner Res 2:595–610.
737. Doube M, Firth EC, Boyde A 2007 Variations in articular
calcified cartilage by site and exercise in the 18-month-old
equine distal metacarpal condyle. Osteoarthritis Cartilage
15:1283–1292.
738. Barnard JC, Williams AJ, Rabier B, Chassande O, Samarut J,
Cheng SY, Bassett JH, Williams GR 2005 Thyroid hormones
regulate fibroblast growth factor receptor signaling during
chondrogenesis. Endocrinology 146:5568–5580.
739. Stevens DA, Harvey CB, Scott AJ, O’Shea PJ, Barnard JC,
Williams AJ, Brady G, Samarut J, Chassande O, Williams
GR 2003 Thyroid hormone activates fibroblast growth
factor receptor-1 in bone. Mol Endocrinol 17:1751–1766.
740. Williams AJ, Robson H, Kester MH, van Leeuwen JP,
Shalet SM, Visser TJ, Williams GR 2008 Iodothyronine
deiodinase enzyme activities in bone. Bone 43:126–134.
741. Rabier B, Williams AJ, Mallein-Gerin F, Williams GR,
Chassande O 2006 Thyroid hormone-stimulated differen-
tiation of primary rib chondrocytes in vitro requires thyroid
hormone receptor beta. J Endocrinol 191:221–228.
742. Robson H, Siebler T, Stevens DA, Shalet SM, Williams GR
2000 Thyroid hormone acts directly on growth plate
chondrocytes to promote hypertrophic differentiation
and inhibit clonal expansion and cell proliferation. En-
docrinology 141:3887–3897.
166 BIANCO ET AL.
743. O’Shea PJ, Guigon CJ, Williams GR, Cheng SY 2007 Regula-
tion of fibroblast growth factor receptor-1 by thyroid hor-
mone: identification of a thyroid hormone response element
in the murine Fgfr1 promoter. Endocrinology 148:5966–5976.
744. [Deleted.]
745. Cheron RG, Kaplan MM, Larsen PR 1979 Physiological and
pharmacological influences on thyroxine to 3,5,3¢- triiodo-
thyronine conversion and nuclear 3,5,3¢-triiodothyronine
binding in rat anterior pituitary. J Clin Invest 64:1402–1414.
746. DeJong M, Docter R, Van der Hoek H, Krenning E, van der
Heide D, Quero C, Plaisier P, Vos R, Hennemann G 1994
Different effects of amiodarone on transport of T
4
and T
3
into the perfused liver. Am J Physiol Endocrinol Metab
266:E44–E49.
747. Wassen FW, Moerings EP, Van Toor H, De Vrey EA, Hen-
nemann G, Everts ME 1996 Effectsofinterleukin-1betaon
thyrotropin secretion and thyroid hormone uptake in cultured
rat anterior pituitary cells. Endocrinology 137:1591–1598.
748. Wassen FW, Moerings EP, van Toor H, Hennemann G,
Everts ME 2000 Thyroid hormone uptake in cultured rat
anterior pituitary cells: effects of energy status and biliru-
bin. J Endocrinol 165:599–606.
749. Everts ME, Visser TJ, Moerings, E.P.C.M., Docter R, van
Toor H, Tempelaars, A.M.P., DeJong M, Krenning EP,
Hennemann G 1994 Uptake of triiodothyroacetic acid and
its effect on thyrotropin secretion in cultured anterior pi-
tuitary cells. Endocrinology 135:2700–2707.
750. Sato K, Han DC, Fujii Y, Tsushima T, Shizume K 1987
Thyroid hormone stimulates alkaline phosphatase activity
in cultured rat osteoblastic cells (ROS 17/2.8) through
3,5,3¢-triiodo-L-thyronine nuclear receptors. Endocrinology
120:1873–1881.
751. Sato Y, Nakamura R, Satoh M, Fujishita K, Mori S, Ishida S,
Yamaguchi T, Inoue K, Nagao T, Ohno Y 2005 Thyroid
hormone targets matrix Gla protein gene associated with
vascular smooth muscle calcification. Circ Res 97:550–557.
752. Liu Y, Fu L, Chen DG, Deeb SS 2007 Identification of novel
retinal target genes of thyroid hormone in the human WERI
cells by expression microarray analysis. Vision Res 47:
2314–2326.
Address correspondence to:
Antonio C. Bianco, MD, PhD
Division of Endocrinology, Diabetes and Metabolism
University of Miami Miller School of Medicine
Suite 816 Dominion Towers
1400 NW 10th Ave.
Miami, FL 33136
E-mail: abianco@deiodinase.org
Abbreviations Used
2D ¼ two-dimensional
3D ¼ three-dimensional
5¢D ¼ 5¢-deiodination
5D ¼ 5-deiodination
ATA ¼ American Thyroid Association
BAT ¼ brown adipose tissue
BSA ¼ bovine serum albumin
BSE-SEM ¼ back scattered electron-scanning electron
microscopy
bTSH ¼ bovine TSH
Abbreviations Used (Cont.)
BW ¼ body weight
CAR ¼ constitutive androstane receptor
ChIP ¼ chromatin immunoprecipitation
CNS ¼ central nervous system
CT ¼ computerized tomography
Ct ¼ critical threshold
D1 ¼ type I deiodinase
D2 ¼ type II deiodinase
D3 ¼ type III deiodinase
DITPA ¼ 3,5-diiodothyropropionic acid
DTT ¼ dithiothreitol
DUOX2 ¼ dual oxidase 2
ECG ¼ electrocardiogram
EDL ¼ extensor digitorum longus
EDTA ¼ ethylenediaminetetraacetic acid
ELISA ¼ enzyme-linked immunosorbent assay
EMSA ¼ electromobility shift assays
EYFP ¼ enhanced yellow fluorescent protein
FBS ¼ fetal bovine serum
FT
3
I ¼ free T
3
index
FT
4
I ¼ free T
4
index
GFP ¼ green fluorescent protein
GPD ¼ glycerophosphate dehydrogenase
H&E ¼ hematoxylin and eosin
HFUS ¼ high frequency ultrasound
HPLC ¼ high performance liquid chromatography
HPT ¼ hypothalamus-pituitary-thyroid axis
iBAT ¼ interscapular BAT
ICV ¼ intra-cerebro-ventricular
IHC ¼ immunohistochemistry
IRMA ¼ immune radiometric assay
KClO
4
¼ potassium perchlorate
KO ¼ knock-out
LDL ¼ low-density lipoprotein
LID ¼ low-iodine diet
LPS ¼ lipopolysaccharide
L-T
4
¼ levothyroxine
LV ¼ left ventricle (ventricular)
MCT ¼ monocarboxylate transporter
MHC ¼ myosin heavy chain
MLC ¼ myosin light chain
MMI ¼ methimazole
MRI ¼ magnetic resonance imaging
NaClO
4
¼ sodium perchlorate
NIS ¼ sodium iodine symporter
NTI ¼ nonthyroidal illness
OATP ¼ organic anion transporting polypeptide
PAPS ¼ 3¢-phosphoadenosine-5¢-phosphosulfate
PBS ¼ phosphate-buffered saline
PCR ¼ polymerase chain reaction
PET ¼ positron emission tomography
PTU ¼ 6-n-propyl-2-thiouracil
PXR ¼ pregnane X receptor
RAIU ¼ radioactive iodide uptake
RANKL ¼ receptor activator of nuclear factor
kappa B-ligand
rhTSH ¼ recombinant human TSH
RIA ¼ radioimmunoassay
RNAseq ¼ RNA sequences
RQ ¼ respiratory quotient
rT
3
¼ 3,3¢,5¢-triiodothyronine
INVESTIGATING THYROID HORMONE ECONOMY AND ACTION 167
Abbreviations Used (Cont.)
RT-qPCR ¼ reverse transcriptase quantitative PCR
RV ¼ right ventricle (ventricular)
SR ¼ sarcoplasmic reticulum
SERCA1 and 2 ¼ sarcoplasmic/endoplasmic reticulum
Ca
++
ATPase 1 and 2
SIMS ¼ secondary ion mass spectrometry
SLC26A4/PDS ¼ pendrin
SNS ¼ sympathetic nervous system
SOL ¼ soleus
SPECT ¼ single photon emission computed
tomography
SR ¼ sarcoplasmic reticulum
SULT ¼ sulfotranferase
T
2
¼ 3,3¢-diiodothyronine
T
3
¼ 3,3¢,5-triiodothyr onine
T
3
RE ¼ T
3
responsive element
T
3
S ¼ T
3
sulfate
T
4
¼ thyroxine
TBG ¼ thyroxine binding globulin
Abbreviations Used (Cont.)
TCA ¼ trichloroacetic acid
T/ebp ¼ thyroid-specific enhancer-binding
protein
Tetrac ¼ tetraiodothyroacetic acid
Tm ¼ equilibrium time point
TR ¼ thyroid hormone receptor
TRAcP ¼ tartrate-resistant acid phosphatase
TRH ¼ thyrotropin releasing hormone
TRIAC ¼ tiratricol (3,5,3¢-triiodothyroacetic acid)
Tris-HCl ¼ 2-amino-2-hydroxymethyl-1,
3-propanediol hydrochloride
TSH ¼ thyrotropin
TSHR ¼ TSH receptor
UDPGA ¼ UDP-glucuronic acid
UGT ¼ UDP-glucuronyltransferase
UPLC ¼ ultrahigh performance liquid
chromatography
VO
2
¼ oxygen consumption
VSMC ¼ vascular smooth muscle cell
168 BIANCO ET AL.
This article has been cited by:
1. R Thomas Zoeller, Laura N Vandenberg. 2015. Assessing dose–response relationships for endocrine disrupting chemicals (EDCs):
a focus on non-monotonicity. Environmental Health 14. . [CrossRef]
2. Tatiana L. Fonseca, Gustavo W. Fernandes, Elizabeth A. McAninch, Barbara M. L. C. Bocco, Sherine M. Abdalla, Miriam
O. Ribeiro, Petra Mohácsik, Csaba Fekete, Daofeng Li, Xiaoyun Xing, Ting Wang, Balázs Gereben, Antonio C. Bianco. 2015.
Perinatal deiodinase 2 expression in hepatocytes defines epigenetic susceptibility to liver steatosis and obesity. Proceedings of the
National Academy of Sciences 112, 14018-14023. [CrossRef]
3. Laurberg Peter, Andersen Stine Linding. 2015. Antithyroid Drug Use in Pregnancy and Birth Defects: Why Some Studies Find
Clear Associations, and Some Studies Report None. Thyroid 25:11, 1185-1190. [Abstract] [Full Text HTML] [Full Text PDF]
[Full Text PDF with Links]
4. Raymond M. David, Randy D. White, Michael J. Larson, Jay K. Herman, Rainer Otter. 2015. Toxicity of Hexamoll® DINCH®
following intravenous administration. Toxicology Letters 238, 100-109. [CrossRef]
5. Balázs Gereben, Elizabeth A. McAninch, Miriam O. Ribeiro, Antonio C. Bianco. 2015. Scope and limitations of iodothyronine
deiodinases in hypothyroidism. Nature Reviews Endocrinology 11, 642-652. [CrossRef]
6. Patricia Joseph-Bravo, Lorraine Jaimes-Hoy, Rosa María Uribe, Jean-Louis Charli. 2015. 60 YEARS OF
NEUROENDOCRINOLOGY: TRH, the first hypophysiotropic releasing hormone isolated: control of the pituitary–thyroid
axis. Journal of Endocrinology 226, T85-T100. [CrossRef]
7. S. W. Jansen, A. A. Akintola, F. Roelfsema, E. van der Spoel, C. M. Cobbaert, B. E. Ballieux, P. Egri, Z. Kvarta-Papp, B. Gereben,
C. Fekete, P. E. Slagboom, J. van der Grond, B. A. Demeneix, H. Pijl, R. G. J. Westendorp, D. van Heemst. 2015. Human
longevity is characterised by high thyroid stimulating hormone secretion without altered energy metabolism. Scientific Reports
5, 11525. [CrossRef]
8. Simon C. Roberts, Antonio C. Bianco, Heather M. Stapleton. 2015. Disruption of Type 2 Iodothyronine Deiodinase Activity in
Cultured Human Glial Cells by Polybrominated Diphenyl Ethers. Chemical Research in Toxicology 28, 1265-1274. [CrossRef]
9. Joao Pedro Werneck de Castro, Tatiana L. Fonseca, Cintia B. Ueta, Elizabeth A. McAninch, Sherine Abdalla, Gabor Wittmann,
Ronald M. Lechan, Balazs Gereben, Antonio C. Bianco. 2015. Differences in hypothalamic type 2 deiodinase ubiquitination
explain localized sensitivity to thyroxine. Journal of Clinical Investigation 125, 769-781. [CrossRef]
10. Jonklaas Jacqueline, Bianco Antonio C., Bauer Andrew J., Burman Kenneth D., Cappola Anne R., Celi Francesco S., Cooper David
S., Kim Brian W., Peeters Robin P., Rosenthal M. Sara, Sawka Anna M.. 2014. Guidelines for the Treatment of Hypothyroidism:
Prepared by the American Thyroid Association Task Force on Thyroid Hormone Replacement. Thyroid 24:12, 1670-1751.
[Abstract] [Full Text HTML] [Full Text PDF] [Full Text PDF with Links]
11. Sherine M. Abdalla, Antonio C. Bianco. 2014. Defending plasma T3 is a biological priority. Clinical Endocrinology 81:10.1111/
cen.2014.81.issue-5, 633-641. [CrossRef]
12. Rob Janssen, Marian J. Zuidwijk, Diederik W. D. Kuster, Alice Muller, Warner S. Simonides. 2014. Thyroid Hormone-Regulated
Cardiac microRNAs are Predicted to Suppress Pathological Hypertrophic Signaling. Frontiers in Endocrinology 5. . [CrossRef]
13. P. Egri, B. Gereben. 2014. Minimal requirements for ubiquitination-mediated regulation of thyroid hormone activation. Journal
of Molecular Endocrinology 53, 217-226. [CrossRef]
14. G. W. Fernandes, C. B. Ueta, T. L. Fonseca, C. H. A. Gouveia, C. L. Lancellotti, P. C. Brum, M. A. Christoffolete, A. C. Bianco,
M. O. Ribeiro. 2014. Inactivation of the adrenergic receptor 2 disrupts glucose homeostasis in mice. Journal of Endocrinology
221, 381-390. [CrossRef]
15. T. L. Fonseca, J. P. Werneck-De-Castro, M. Castillo, B. M. L. C. Bocco, G. W. Fernandes, E. A. McAninch, D. L. Ignacio, C. C.
S. Moises, A. R. Ferreira, B. Gereben, A. C. Bianco. 2014. Tissue-Specific Inactivation of Type 2 Deiodinase Reveals Multilevel
Control of Fatty Acid Oxidation by Thyroid Hormone in the Mouse. Diabetes 63, 1594-1604. [CrossRef]
... The other cause is errors of TH genesis (dyshormonogenesis). The most common CH rodent models are the drug-induced models with anti-thyroid drugs such as propylthiouracil or 2-mercapto-1-methylimidazole (MMI) [53,54]. Nicholson and Altman reported a reduced cerebellar weight; a prolonged cell proliferation in the external granule cell layer (EGL) and retarded EGL disappearance; a retarded cell differentiation in the molecular and internal granule cell layers, and terminal increases in the numbers of granule cells and astrocytes; and a decrease in the numbers of basket cells. ...
... This indicates that RORα may be involved in TH-regulated gene expression in the developing cerebellum. In fact, TH regulates RORα expression during the first two postnatal weeks [54,66], indicating that TH may alter gene expression critical for cerebellar development through RORα regulation. Furthermore, RORα augments TR-mediated transcription, whereas staggerer-type mutant RORα does not have such action [108]. ...
Article
Full-text available
The proper organized expression of specific genes in time and space is responsible for the organogenesis of the central nervous system including the cerebellum. The epigenetic regulation of gene expression is tightly regulated by an intrinsic intracellular genetic program, local stimuli such as synaptic inputs and trophic factors, and peripheral stimuli from outside of the brain including hormones. Some hormone receptors are expressed in the cerebellum. Thyroid hormones (THs), among numerous circulating hormones, are well-known major regulators of cerebellar development. In both rodents and human, hypothyroidism during the postnatal developmental period results in abnormal morphogenesis or altered function. THs bind to the thyroid hormone receptors (TRs) in the nuclei and with the help of transcriptional cofactors regulate the transcription of target genes. Gene regulation by TR induces cell proliferation, migration, and differentiation, which are necessary for brain development and plasticity. Thus, the lack of TH action mediators may directly cause aberrant cerebellar development. Various kinds of animal models have been established in a bid to study the mechanism of TH action in the cerebellum. Interestingly, the phenotypes differ greatly depending on the models. Herein we summarize the actions of TH and TR particularly in the developing cerebellum.
... As a core clock protein, functional disruptions in RORα can directly impair the regulation of circadian rhythms, thereby increasing the risk for developing depressive disorders [47][48][49]. Moreover, RORα participates in the regulation of neurogenesis and neuroplasticity through modulating the expression of crucial molecules such as BDNF and 5-HT, processes that are intimately linked to the neurobiological underpinnings of MDD [50,51]. ...
... Elevated BDNF levels may be crucial for reversing motor and cognitive impairments in neurodegenerative disorders (Kumar et al. 2013). According to Several research, thyroxin increases BDNF level in developing neurons (Koibuchi et al. 2001;Camboni et al. 2003). Studies have revealed that basal levels of BDNF protein evaluated by Western blot, in addition to the quantity of the signaling molecules implicated in synaptic plasticity (e.g., calcineurin, CREB, MAP kinase) are reduced in rats following surgical ablation of the thyroid gland (Gilbert and Lasley 2013). ...
Article
Full-text available
Huntington’s disease (HD) is a neurodegenerative disorder characterized by cognitive deficits and motor function. Levothyroxine (L-T4) is a synthetic form of Thyroxine (T4), which can improve cognitive ability. The aim of the present study was to determine the neuroprotective effect of L-T4 administration in rats with 3-nitropropionic acid (3-NP)-induced Huntingtonʼs disease. Forty-eight Wistar male rats were divided into six groups (n = 8): Group 1 control group that received physiological saline, Group 2 and 3: which received L-T4 (30 and 100 μg/kg), Group 4: HD group that received 3-NP and Groups 5 and 6: The treatment of the HD rats with L-T4 (30 and 100 μg/kg). Spatial memory, locomotor activity, and frequency of neuronal firing were assessed. After decapitation, the Brain-Derived Neurotrophic Factor (BDNF) and Total antioxidant capacity (TAC) levels in the striatum was measured. The results showed that the indices of spatial memory (mean path length and latency time) and motor dysfunction (immobility time) significantly increased, while time spent in the goal quadrant, swimming speed, spike rate, and striatum levels of BDNF significantly decreased in the HD group compared to the control group. L-T4 treatment significantly enhanced time spent in the goal quadrant, swimming speed, motor activity (number of line crossing and rearing), spike rate and striatal BDNF level. This research showed that L-T4 prevented the disruption of motor activity and cognitive deficiencies induced by 3-NP. The beneficial effects of L-T4 may be due to an increase in the concentration of BDNF and enhancement of the spike rate in the striatum. Graphical abstract
... In animal models, rats with congenital hypothyroidism have severe impairments of motor coordination and cerebellar morphogenesis, characterized by small somata, poor dendritic arborization of Purkinje cells (PCs), and retarded migration of granule cells (GCs) during the first 2 postnatal weeks (4). Although the expressions of several important genes for cerebellar function are altered in mice with congenital hypothyroidism at postnatal day (PD) 15, expressions of most genes are restored later in adulthood (5,6). Nevertheless, motor deficits persist in adult mice with congenital hypothyroidism (7). ...
Article
Full-text available
Thyroid hormones (THs) regulate gene expression by binding to nuclear TH receptors (TRs) in the cell. THs are indispensable for brain development. However, we have little knowledge about how congenital hypothyroidism in neurons affects functions of the central nervous system in adulthood. Here, we report specific TH effects on functional development of the cerebellum by using transgenic mice overexpressing a dominant-negative TR (Mf-1) specifically in cerebellar Purkinje cells (PCs). Adult Mf-1 mice displayed impairments in motor coordination and motor learning. Surprisingly, long-term depression (LTD)–inductive stimulation caused long-term potentiation (LTP) at parallel fiber (PF)–PC synapses in adult Mf-1 mice, although there was no abnormality in morphology or basal properties of PF–PC synapses. The LTP phenotype was turned to LTD in Mf-1 mice when the inductive stimulation was applied in an extracellular high-Ca ²⁺ condition. Confocal calcium imaging revealed that dendritic Ca ²⁺ elevation evoked by LTD-inductive stimulation is significantly reduced in Mf-1 PCs but not by PC depolarization only. Single PC messenger RNA quantitative analysis showed reduced expression of SERCA2 and IP 3 receptor type 1 in Mf-1 PCs, which are essential for mGluR1-mediated internal calcium release from endoplasmic reticulum in cerebellar PCs. These abnormal changes were not observed in adult-onset PC-specific TH deficiency mice created by adeno-associated virus vectors. Thus, we propose the importance of TH action during neural development in establishing proper cerebellar function in adulthood, independent of its morphology. The present study gives insight into the cellular and molecular mechanisms underlying congenital hypothyroidism–induced dysfunctions of central nervous system and cerebellum.
... The vehicle control group was injected with the corresponding amount of DMSO, since RGFP966 was dissolved in DMSO. The last group was administered 20 µg/kg BW of T 4 (Merck, Darmstadt, Germany), which is the amount used to rescue PTU-induced hypothyroidism in perinatal mice [47]. The T 4 group was also injected with DMSO because the effect of DMSO should not be ignored during organogenesis. ...
Article
Full-text available
Perinatal hypothyroidism impairs cerebellar organogenesis and results in motor coordination defects. The thyroid hormone receptor binds to corepressor complexes containing histone deacetylase (HDAC) 3 in the absence of ligands and acts as a transcriptional repressor. Although histone acetylation status is strongly correlated with transcriptional regulation, its role in cerebellar development remains largely unknown. We aimed to study whether the cerebellar developmental defects induced by perinatal hypothyroidism can be rescued by treatment with a specific HDAC3 inhibitor, RGFP966. Motor coordination was analyzed using three behavioral tests. The cerebella were subjected to RT-qPCR and chromatin immunoprecipitation assays for acetylated histone H3. The treatment with RGFP966 partially reversed the cerebellar morphological defects in perinatal hypothyroid mice. These findings were associated with the alleviation of motor coordination defects in these mice. In addition, the RGFP966 administration increased the mRNA levels of cerebellar thyroid hormone-responsive genes. These increases were accompanied by augmented histone acetylation status at these gene loci. These findings indicate that HDAC3 plays an important role in the cerebellar developmental defects induced by perinatal hypothyroidism. The HDAC3 inhibitor might serve as a novel therapeutic agent for hypothyroidism-induced cerebellar defects by acetylating histone tails and stimulating transcription at thyroid hormone-responsive gene loci.
... The majority of cell subsets in the developing cerebellum express the receptors for this hormone, thyroid receptors TRs [58]. Activation of these receptors leads to activation of different signaling pathways related to proliferation and differentiation processes [59,60], such EGF pathway, as it will be discussed later in this review. ...
Article
Full-text available
Astrocytes, initially described as merely support cells, are now known as a heterogeneous population of cells actively involved in a variety of biological functions such as: neuronal migration and differentiation; regulation of cerebral blood flow; metabolic control of extracellular potassium concentration; and modulation of synapse formation and elimination; among others. Cerebellar glial cells have been shown to play a significant role in proliferation, differentiation, migration, and synaptogenesis. However, less evidence is available about the role of neuron-astrocyte interactions during cerebellar development and their impact on diseases of the cerebellum. In this review, we will focus on the mechanisms underlying cellular interactions, specifically neuron-astrocyte interactions, during cerebellar development, function, and disease. We will discuss how cerebellar glia, astrocytes, and Bergmann glia play a fundamental role in several steps of cerebellar development, such as granule cell migration, axonal growth, neuronal differentiation, and synapse formation, and in diseases associated with the cerebellum. We will focus on how astrocytes and thyroid hormones impact cerebellar development. Furthermore, we will provide evidence of how growth factors secreted by glial cells, such as epidermal growth factor and transforming growth factors, control cerebellar organogenesis. Finally, we will argue that glia are a key mediator of cerebellar development and that identification of molecules and pathways involved in neuron-glia interactions may contribute to a better understanding of cerebellar development and associated disorders.
... RORα-deficient mice do not completely lack opsin expression, but S-and M-opsins are significantly downregulated 52 . Moreover, expression of the RORα gene may be regulated by TH, because expression levels were partly downregulated in brains of hypothyroid mice compared to T4-replaced controls 53 . Taken together, the naturally hypothyroid mole-rat retina might not only lead to low M-opsin expression levels, but rather to a general downregulation of both cone opsins (and other cone-specific genes) due to missing RORα signalling. ...
Article
Full-text available
Mammals usually possess a majority of medium-wavelength sensitive (M-) and a minority of short-wavelength sensitive (S-) opsins in the retina, enabling dichromatic vision. Unexpectedly, subterranean rodents from the genus Fukomys exhibit an S-opsin majority, which is exceptional among mammals, albeit with no apparent adaptive value. Because thyroid hormones (THs) are pivotal for M-opsin expression and metabolic rate regulation, we have, for the first time, manipulated TH levels in the Ansell’s mole-rat (Fukomys anselli) using osmotic pumps. In Ansell’s mole-rats, the TH thyroxine (T4) is naturally low, likely as an adaptation to the harsh subterranean ecological conditions by keeping resting metabolic rate (RMR) low. We measured gene expression levels in the eye, RMR, and body mass (BM) in TH-treated animals. T4 treatment increased both, S- and M-opsin expression, albeit M-opsin expression at a higher degree. However, this plasticity was only given in animals up to approximately 2.5 years. Mass-specific RMR was not affected following T4 treatment, although BM decreased. Furthermore, the T4 inactivation rate is naturally higher in F. anselli compared to laboratory rodents. This is the first experimental evidence that the S-opsin majority in Ansell’s mole-rats is a side effect of low T4, which is downregulated to keep RMR low.
Article
Full-text available
Thyroid hormone exerts marked effects on cardiovascular function. Expression of cardiac alpha- and beta-myosin heavy chain (MHC) isoforms can be altered in response to thyroid hormone as well as by hemodynamic changes imposed on the heart. The molecular mechanisms that mediate these changes are not completely known. We studied the contractile and thyroid hormone responsiveness of the betaMHC promoter in both cultured cardiac myocytes and in vivo by direct DNA transfer. Using transient transfection of neonatal rat cardiomyocytes, the activities of recombinant reporter plasmids containing betaMHC 5'-flanking sequences terminating at positions -2250, -1145, -670, and -354 were decreased significantly in cultures containing L-T3 (50 nM). Similar deletion analysis showed that 5'-flanking regions terminating within -2250 to -151 bp were contractility responsive; however, deletion to position -126 attenuated this response. In vivo betaMHC promoter activity, determined by injecting the recombinant plasmid into th...
Article
Recent evidence suggests that neuropeptide Y (NPY), originating in neurons in the hypothalamic arcuate nucleus, is an important mediator of the effects of leptin on the central nervous system. As these NPY neurons innervate hypophysiotropic neurons in the hypothalamic paraventricular nucleus (PVN) that produce the tripeptide, TRH, we raised the possibility that NPY may be responsible for resetting of the hypothalamic-pituitary-thyroid (HPT) axis during fasting. To test this hypothesis, the effects of intracerebroventricularly administered NPY on circulating thyroid hormone levels and proTRH messenger RNA in the PVN were studied by RIA and in situ hybridization histochemistry, respectively. NPY administration suppressed circulating levels of thyroid hormone (T3 and T4) and resulted in an inappropriately normal or low TSH. These alterations were associated with a significant suppression of proTRH messenger RNA in the PVN, indicating that NPY infusion had resulted in a state of central hypothyroidism. Simila...
Article
beta 1 and beta 2 are functional thyroid hormone receptors (TRs) that are generated from the same genomic locus by splicing of a different amino terminus onto a common carboxyl region containing the DNA and hormone binding domains. TR beta 1 is widely expressed whereas TR beta 2 is found primarily in the pituitary gland although low levels of expression have been described in other tissues. To gain insight into the mechanisms governing expression of this complex transcriptional unit, we cloned mouse genomic fragments containing the common carboxyl terminus as well as the unique TR beta 2 amino-terminal sequence that was located at least 25 kilobases upstream. The DNA and ligand binding exons are identical in size and location of their boundaries to those of the human TR beta 1 gene. To determine whether the region 5' of the TR beta 2 amino terminus represented the promoter region, we examined it for sites of transcriptional initiation and for its ability to function as a promoter in TR beta 2-expressing t...
Article
The effects of interleukin-1 beta (IL-1 beta) and tumor necrosis factor-alpha (TNF alpha) on basal and TRH-induced TSH release, and the effects of IL-1 beta on the uptake of [125I]T3 and [125I]T4 and on nuclear binding of [125I]T3 were examined. Furthermore, the release of other anterior pituitary hormones in the presence of IL-1 beta was measured. Anterior pituitary cells from male Wistar rats were cultured for 3 days in medium containing 10% FCS. Incubation were performed at 37 C in medium with 0.5% BSA for measurement of [125I]T3 uptake and with 0.1% BSA for measurement of [125I]T4 uptake. Exposure to IL-1 beta (1 pM-1 nM) or TNF alpha (100 pM) for 2-4 h resulted in a significant decline in TSH release, which was almost 50% (P < 0.05) for 1 nM IL-1 beta and 24% (P < 0.05) for 100 pM TNF alpha. Measurement of other anterior pituitary hormones (FSH, LH, PRL, and ACTH) in the same incubation medium showed that IL-1 beta did not alter their release. When the effects of IL-1 beta (1 pM-1 nM) and TNF alpha (...
Article
Human type II iodothyronine deiodinase (D2) catalyzes the activation of T4 to T3. The D2 enzyme, like the type I (D1) and type III (D3) deiodinases, contains a selenocysteine (SeC) residue (residue 133 in D2) in the highly conserved catalytic center. Remarkably, all of the D2 proteins cloned so far have an alanine two residue-amino terminal to the SeC, whereas all D1 and D3 proteins contain a cysteine at this position. A cysteine residue in the catalytic center could assist in enzymatic action by providing a nucleophilic sulfide or by participating in redox reactions with a cofactor or enzyme residues. We have investigated whether D2 mutants with a cysteine (A131C) or serine (A131S) two-residue amino terminal to the SeC are enzymatically active and have characterized these mutants with regard to substrate affinity, reducing cofactor interaction and inhibitor profile. COS cells were transfected with expression vectors encoding wild-type (wt) D2, D2 A131C, or D2 A131S proteins. Kinetic analysis was performed on homogenates with dithiothreitol (DTT) as reducing cofactor. The D2 A131C and A131S mutants displayed similar Michaelis-Menten constant values for T4 (5 nM) and reverse T3 (9 nM) as the wt D2 enzyme. The limiting Michaelis-Menten constant for DTT of the D2 A131C enzyme was 3-fold lower than that of the wt D2 enzyme. The wt and mutant D2 enzymes are essentially insensitive to propylthiouracil [concentration inhibiting 50% of activity (IC50) > 2m M] in the presence of 20 mM DTT, but when tested in the presence of 0.2 mM DTT the IC50 value for propylthiouracil is reduced to about 0.1 mM. During incubations of intact COS cells expressing wt D2, D2 A131C, or D2 A131S, addition of increasing amounts of unlabeled T4 resulted in the saturation of [ 125 I]T4 deiodination, as reflected in a decrease of [ 125 I]T3 release into the medium. Saturation first appeared at medium T4 concentrations between 1 and 10 nM. In conclusion: substitution of cysteine for a conserved alanine residue in the catalytic center of the D2 protein does not inactivate the enzyme in vitro and in situ, but rather improves the interaction with the reducing cofactor DTT in vitro .( Endocrinology 143: 1190 –1198, 2002)
Article
Type I iodothyronine 5' deiodinase (5'DI) contains selenocysteine, encoded by a UGA codon, and this amino acid is essential for maximum catalytic efficiency in this enzyme. We recently showed that translation of UGA as selenocysteine in this protein requires a specific sequence of about 250 nucleotides in the 3' untranslated region of the messenger RNA. Translation of a 5'DI cysteine mutant does not require the 3' untranslated region. To examine both the efficiency of UGA codon recognition and the relative catalytic efficiency of selenocysteine vs. cysteine in 5'DI, we used bromoacetyl 125I-T3 labeling to quantitate transiently expressed selenocysteine (wild type) and cysteine containing type I iodothyronine deiodinases in transfected COS-7 and JEG-3 cell lines. Kinetic analyses of the same cell sonicates were performed to determine the apparent maximum velocity and Michaelis-Menten constant values for reverse T3 5' deiodination. COS-7 cells express the cysteine mutant protein at about 20-fold and JEG-3 cells about 400-fold higher levels than the selenoenzyme. However, in both cell types, the apparent catalytic constant values were at least 100-fold higher for the wild-type enzyme, compared with the cysteine mutant. These results indicate that cell lines differ markedly in their capacity to translate UGA-containing messenger RNAs. The much higher catalytic constant values for the selenium-containing enzyme illustrate the biochemical advantage of this element as compared with sulfur in the catalysis of iodothyronine deiodination.
Article
There are three main situations in which an association has been noted between anomalies of thyroid function and mental retardation: (A) Severe endemic goiter associated since the XVIth century with the birth of deaf-mute and imbecile inhabitants, known as cretins (1). This association was recognized before it was even known that goiter is an enlarged thyroid. (B) Congenital hypothyroidism associated with severe mental retardation since Curling described two cases in the mid-XIXth century (2). (C) Maternal hypothyroxinemia, with or without clinical hypothyroidism, associated since the studies of Man et al. (3) with decreased mean I.Q. of the progeny.