Exp Clin Endocrinol Diabetes 2022; 130(05): 296-302
DOI: 10.1055/a-1716-7980
Article

What is the Role of Thyroid Hormone Receptor Alpha 2 (TRα2) in Human Physiology?

Sarah Paisdzior
1   Institute of Experimental Pediatric Endocrinology, Charité – Universitätsmedizin Berlin, corporate member of Freie Universität Berlin and Humboldt-Universität zu Berlin, Berlin, Germany
,
Markus Schuelke
2   NeuroCure Cluster of Excellence; Charité – Universitätsmedizin Berlin, corporate member of Freie Universität Berlin and Humboldt-Universität zu Berlin, Berlin, Germany
3   Department of Neuropediatrics, Charité – Universitätsmedizin Berlin, corporate member of Freie Universität Berlin and Humboldt-Universität zu Berlin, Berlin, Germany
,
Heiko Krude
1   Institute of Experimental Pediatric Endocrinology, Charité – Universitätsmedizin Berlin, corporate member of Freie Universität Berlin and Humboldt-Universität zu Berlin, Berlin, Germany
› Author Affiliations
 

Abstract

Thyroid hormone receptors are nuclear receptors that function as transcription factors and are regulated by thyroid hormones. To date, a number of variants and isoforms are known. This review focuses on the thyroid hormone receptor α (TRα), in particular TRα2, an isoform that arises from alternative splicing of the THRA mRNA transcript. Unlike the TRα1 isoform, which can bind T3, the TRα2 isoform lacks a ligand-binding domain but still binds to DNA thereby antagonizing the transcriptional activity of TRα1. Although a regulatory role has been proposed, the physiological function of this TRα2 antagonism is still unclear due to limited in vitro and mouse model data. Recently, the first patients with resistance to thyroid hormone due to mutations in THRA, the TRα encoding gene, affecting the antagonistic function of TRα2 were described, suggesting a significant role of this particular isoform in human physiology.


#

The thyroid hormones (TH) triiodothyronine (T3) and thyroxine (T4) are important regulators of biological functions. Lack of TH action can lead to developmental, metabolic, and cardiovascular diseases. At the cellular level, nuclear receptors called “thyroid hormone receptors” (THRs) mediate TH function [1]. Heritable syndromes of impaired TH sensitivity include, among others, resistance to thyroid hormone (RTH), which is caused by mutations in genes encoding for THRs, in particular the two genes THRA and THRB [2]. In general, patients’ symptoms depend on the expression pattern of the affected gene and the resulting functional defect. Refetoff et al. coined the acronym RTH when they described the first patient in 1967 [3], who, 20 years later, was found to have a homozygous deletion in the THRB gene encoding TRβ [4]. During this time, additional patients with RTH were identified, most of whom carried a heterozygous missense mutation in THRB [5] [6]. Patients with RTH have elevated T3 and T4 serum concentrations, with some having elevated, but never suppressed thyrotropin (TSH) levels, making the association with TH-dependent disease relatively clear. The phenotype is variable, and few patients show severe symptoms such as attention-deficit hyperactivity disorder (ADHD), tachycardia, or goiter [7] [8]. TRβ is mainly responsible for the negative feedback loop regulating the hypothalamus-pituitary-thyroid axis [9] [10], which explains why TSH is not suppressed despite high T3 and T4 levels.

The two genes encoding the TH receptors are TRα (THRA) and TRβ (THRB), therefore, the differences in the phenotype of patients with TH receptor gene mutations are likely due to different expression patterns of the two isoforms. With the development of new technologies such as whole-exome sequencing (WES), the first THRA mutation was identified in 2012 [11], followed shortly by a second case in the same year identified by Sanger sequencing [12]. Since then, the number of THRA missense mutations has been steadily increasing. These patients lack the changes in blood TH concentration that make other conditions with RTH so distinct, as the hypothalamus-pituitary-thyroid axis is only regulated by TRβ, whereas TRα plays no role in this feedback loop. Therefore, the T3 and T4 serum levels are mainly normal in these patients. It appears that the T3 levels are in the higher normal range while T4 is rather low-normal resulting in a shifted T3/T4 ratio; however, so far, no reference values are available for the T3/4 ratio leaving the relevance of this shift open. The normal T3 and T4 values in most cases make the phenotype even more striking, as THRA mutation carriers have an even more severe phenotype than TRHB mutation carriers with mainly hypothyroid symptoms. These include growth retardation, mild to moderate mental retardation, mild skeletal dysplasia, severe constipation, broad facial features, and bradycardia. Interestingly, most of the identified mutations result in partial or complete loss of function that inhibits gene regulation in a dominant-negative manner [11] [12] [13] [14] [15] [16] [17] [18] [19] [20] [21] [22].

Recently, we described a novel heterozygous point mutation that enhances the function of both TRα isoforms, TRα1 and TRα2 leading to increased T3-activation of TRα1 and increased antagonism of TRα2 [23]. This novel mechanism in RTH due to THRA mutations brings the TRα2 splice variant into focus and raises further interest in its physiological function.

The TRα encoding gene THRA (17q21.1) was previously described as a proto-oncogene “c-erb-A” and was isolated from embryonic chicken, human placental, and rat brain libraries [24] [25]. This gene encodes proteins that share structural features with other nuclear receptors, consisting of a regulatory A/B-domain, a DNA-binding domain (DBD), a hinge region, and a ligand-binding domain (LBD) ([Fig. 1a] ) [26] [27]. C-erb-A was identified as TRα1 by nuclear localization, ability to specifically bind TH and transcriptional regulation of TH-responsive genes [24] [25]. Briefly, TRα1 can interact with TH responsive elements (TRE) located in the promoter regions of TH-regulated genes via the two C4-zinc fingers in the DBD and regulate transcription ([Fig. 2a] ) [28]. Even in an unliganded state, TRs occupy TREs in the function of transcriptional regulators [29].

Zoom Image
Fig. 1 (a) Functional domains of TRα1 and TRα2 isoforms. Published disease-causing variants are marked on the protein domain structure. (b) TRα isoforms resulting from alternative splicing, or different transcription start points, leading to proteins of different molecular weight. (A/B: a regulatory domain; DBD: DNA-binding domain; LBD: ligand-binding domain)
Zoom Image
Fig. 2 (a) The DNA binding domains (DBD) of the unliganded TRα1 homodimer bind to responsive elements of the thyroid hormone receptor (TRE) in the promoter region of target genes. The ligand-binding domains (LBD) are in complex with co-repressors (CoRs) and target gene expression is inhibited. (b) Upon binding of thyroid hormone T3, CoRs are exchanged for co-activators (CoA), and the target genes are then expressed. (c) The weak antagonistic effect of TRα2 is probably due to competition for TREs between both isoforms without interaction with CoRs.

TREs typically consist of a 5´-AGGTCA-3´motif arranged in repeats, either as palindromic, inverted palindromic, or direct repeats spaced by four nucleotides (DR4) [30]. TRs can be positive or negative regulators being able to initiate or inhibit transcription depending on which TRE is bound [31]. The C-terminal LBD is crucial not only for regulating activity through ligand binding but also for interacting with cofactors and dimerizing with other nuclear receptors. Upon binding to T3, the LBD undergoes conformational changes that lead to the replacement of co-repressors (CoRs) by co-activators (CoAs) ([Fig. 2b]) [32]. TRs are known to exist as monomers, but can also form homodimers [33] as well as heterodimers with other nuclear receptors, the most common being the retinoid X receptor (RXR) [34] [35]. Interestingly, RXR has been shown to significantly inhibit transcriptional activity in vitro, suggesting a regulatory role for this heterodimer [36].

In addition to the canonical function as transcription regulators, non-canonical TH signaling has a more rapid effect on the target cell. Flamant et al. proposed a classification of TH action into four subtypes [37]: Type 1 corresponds to the canonical model of TR as a transcription factor by direct binding to DNA, as described above, with TH signaling in mitochondria via the shorter isoform (as described below) also belonging to this type. Type 2 includes signaling via indirect binding to DNA, e. g., by binding to other transcription factors. Type 3 includes signaling independent of DNA binding, such as direct activation of the phosphoinositide 3-kinase/protein kinase B (PI3K/AKT) pathway, which is described in particular for the plasma membrane-bound isoform p30 (see below). Type 4 summarizes TH signaling independent of TRs, e. g., integrin αVβ3, which has been proposed as a membrane receptor for T3 and T4.

Shortly after the discovery of TRα1, the TRα2- isoform was identified resulting from an alternative splice site at exon 9 and transcription of an additional exon 10 [38] [39] ([Fig. 1b]). Interestingly, this isoform is unable to bind TH due to the extended C-terminal LBD [40] leading to an antagonistic effect on TRα1 and TRβ. However, TRα2 has only a weak antagonistic effect on TRα1 and TRβ, as it needs high expression levels to inhibit transcriptional activity ([Fig. 2c]) [41] [42] [43] [44] suggesting a rather minor physiological role for TRα2. Several mechanisms have been suggested to explain this phenomenon: (i) competition for TRE binding sites, (ii) interaction with RXR as the preferred dimerization partner of TRα1 on specific TREs such as DR4 [45], or (iii) a DNA-independent mechanism such as interaction with basal transcriptional factors [41] [46]. Moreover, the phosphorylation state of the elongated C-tail has been shown to influence the antagonistic effect of TRα2 [47]. Interestingly, the inhibitory effect seems to occur only on positive TREs [48], and not on negative TREs. The rather weak effect can be explained by the lack of interaction between TRα2 and CoRs [42] [43].

Nevertheless, protein studies in mouse models and post mortem human brains indicate a high expression ratio of TRα2 to TRα1 [49] [50]. This suggests a T3-independent regulatory role for the physiology of TRα2. Important for this role is that TRα2 has a functional DBD, which still binds to TREs as homo- or heterodimer with RXR, albeit with lower affinity as compared to TRα1 [42] [45]. Therefore, as long as TRα2 is highly expressed, it forms homodimers and occupies TREs without the capability to be activated by T3 and thus acts as an antagonist to TRα1. A direct heterodimer with TRα1, which would result in an even more direct TRα1-antagonism, was not observed on any tested TRE. However, since most dimerization studies of TRs examined the formation of dimers on TREs indirectly by using DNA mobility shift assays, it cannot be excluded that heterodimerization between TRα1 and TRα2 may occur independently of DNA-binding, which was suggested by Katz et al. 1995 [47].

Over time, other TRα isoforms were discovered ([Fig. 1b]), including a suspected third splicing variant, TRα3 that originates from another splicing event in exon 9. Similar to TRα2, this isoform presumably has an elongated C-tail encoded by the sequence of exon 10 (448 amino acids), which is also unable to bind to TH. As only one study was able to detect this isoform, [38] it is vastly understudied, but the predicted structure anticipates the same function as TRα2. Yet another alternative splicing event is responsible for the TRα1-ΔE6 isoform that carries an exchange of exon 6 for the micro-exon 6b. The resulting protein lacks T3-binding capacity but can reduce the transcription-enhancing activity of TRα1. Its proposed role is also regulatory for TRα2 and seems to be important for myocardial development, though the expression pattern of TRα1-ΔE6 suggests additional roles in other tissues [51]. Although it has not been investigated, a TRα2-ΔE6 isoform likely exists as well.

Further truncated isoforms, ΔTRα1 and ΔTRα2, have been observed, which are the result of transcription from an internal promoter in intron 7 and thus are missing the N-terminal domain, but otherwise resemble TRα1 and TRα2 [52]. Additionally, alternative translational start points on the TRα1 mRNA can result in shorter isoforms (p43, p30, p33, and p28) that can be activated by T3 but lack the ability to bind to DNA. They are thought to be bound to the plasma membrane (p30 and p33) [53] [54] or located in the mitochondria (p28 and p43) [55] and maybe responsible for more immediate non-canonical signaling via the MAPK pathway.

Since the first description of a loss-of-function THRA mutation in 2012, about 26 other missense mutations have been identified. One-half of the identified mutations are positioned in the TRα1-specific region of the LBD [11] [12] [14] [16] [19] [55] [56] [57] [58] [59] [60], the other half is located in the coding exons that are shared between TRα1 and TRα2 [13] [15] [16] [17] [18] [19] [20] [21] [22] [61] ( [Fig. 1]). The latter ones could also disturb the function of TRα2, so the phenotypes seen in these patients might provide information about the physiological role of TRα2.

Although several studies in mice have addressed TRα1 and its function [62] [63] [64] [65] [66], the relative contributions of both isoforms to the overall phenotype have proven difficult to dissect. As mentioned earlier, a comparison of protein levels of TR isoforms in mice revealed organ-specific expression, particularly a unique expression pattern of high TRα2 abundance in the central nervous system [49], suggesting an important regulatory role. The specific knockout of TRα2 by adding a strong polyadenylation site that follows the stop codon of TRα1 and transcriptional stop codon resulted in overexpression of TRα1 and a mixed phenotype with hyper- and hypothyroid tissue states [67]. Although this model provides valuable information to the field, it could not fully explain the physiological role of TRα2. Another interesting model is the Pax8 -/- TRα 0/0 compound mouse. Here Pax8, a differentiation factor for thyroid cells was knocked out, which on its own results in the absence of thyroid cells and consequently the complete absence of TH. Without T4 treatment, this defect leads to an early death around weaning time [68]. This model was combined with a TR 0/0 model, harboring a complete deletion of all known TRα isoforms, which on its own was viable, but exhibited reduced growth, delayed bone maturation, moderate hyperthermia, and reduced intestinal mucosal thickness [69]. The Pax8 -/- TRα 0/0 compound model survived without TH-treatment and partially rescued the lethal phenotype of Pax8 -/- mice, but growth was delayed [70]. This study helped to understand how the unliganded receptors might have a physiological function and TH is required to relieve these effects during the postnatal stage. In contrast, Pax8 -/- TRα1 -/- compound models, which still express TRα2 and ΔTRα2 isoforms, have a similar lethal phenotype to Pax8 -/- mice, probably due to an intact TRα2 isoform that could affect the activity of other THRs such TRβ [71]. When comparing these two studies, a possible physiological role for TRα2 is to modulate survival, especially in the first weeks of postnatal development.

Most mutations found in patients were studied in vitro using reporter gene assays based on TRE-dependent luciferase expression and interaction with DNA or with cofactors to show how they inhibit TRα function. For mutations jointly affecting TRα1 and TRα2, some but not all studies have also examined the effects in both splicing isoforms. However, when TRα2 function was tested, most mutations had no measurable effect on antagonistic function. Interestingly, for two mutations (p.A263S and p.N359Y) the inhibitory effect of TRα2 on TRα1 was slightly reduced [15] [16], suggesting a decrease in the dominant-negative effect.

In contrast to all other studies, we recently reported a THRA mutation that resulted in a gain-of-function in both isoforms [23]. A mutated glutamate-to-glycine residue in the first helix of the LBD had a promoting effect on T3-inducible TRα1 activity but also resulted in a gain-of-antagonistic effect for TRα2. Based on the computational model of the LBD, we suspect altered dimerization interphase, although an altered interaction with TRE or cofactors cannot be excluded. Nonetheless, this mutation is the first to enhance TRα2 function by increasing its antagonistic capacity, at least in vitro. At the same time, TRα1 function was increased as well, leading to a pronounced T3 effect. Given this strong gain of function effect of TRα1 in vitro, one would expect a hyperthyroid phenotype of the patients, but this was only the case in patients with mild tachycardia. In fact, we observed more hypothyroid symptoms such as low IQ and global developmental delay, severe constipation, and obesity. Matching these symptoms with our in vitro results suggests that the activated antagonistic effect of the mutant TRα2 was able to counteract the increased activity of the mutant TRα1. Since in most brain regions the TRα2: TRα1 ratio is high [49] [50], the mutant TRα2 appears to significantly suppress the activation of the mutant TRα1, eventually leading to the neuronal hypothyroidism of patients with the THRAp.(E173G) mutation. In other tissues with predominant TRα1 expression, such as cardiomyocytes, the gain-of-function mutation of TRα1 without TRα2 antagonism results in hyperthyroidism-like symptoms. These particular findings of the p.(E173G)-mutant, leading simultaneously to activation of TRα1 and enhanced antagonism of TRα2, suggests that the physiological function of TRα2 is antagonistic to TRα1 function, which appears to be important for the tissue-specific fine-tuning of TH action in target cells.

Overall, 33 years after the discovery of TRα2 and almost 10 years after the first description of patients with mutations in THRA, the first evidence for a physiological effect of TRα2 was found only recently in particular patients carrying a new TRα-mutation. So far, the finding is limited to a single case report and in principle other -potentially genetic- effects can influence the patient’s wide phenotype. However, the obvious antagonistic effect of TRα2, and its increase through this p.(E173G) mutation, proposes a novel mechanism in RTH due to THRA mutations and argues that TRα2 indeed plays a role in controlling the local response of target cells to circulating T3. It is tempting to speculate that any mechanism that increases TRα2 expression relative to TRα1 will decrease the cell response to T3. Moreover, even in tissues with low T3 availability, high levels of TRα2, or mechanisms that increase the DNA-binding of TRα2, are more likely to suppress T3-responsive genes. TRα2 was discovered in the 1980s but few publications on this isoform have appeared in recent decades, thus, it is now time to unravel the physiological role of TRα2 at different developmental time points and in different tissues more thoroughly. Here, special attention must be paid to a clear distinction between the isoforms. Most likely, the potential of single-cell sequencing will stimulate this process and could lead to new and surprising discoveries for the other TRα isoforms that have been little studied so far. Unfortunately, the lack of suitable TRα antibodies, let alone isoform-specific antibodies, prevents the generation of genome-wide chromatin immunoprecipitation sequencing data (ChIP-Seq) of any species. For now, a lot of knowledge regarding TRα isoforms remains to be uncovered.


#

Conflicts of Interest

The authors declare that they had no conflict of interest.


Correspondence

Dr. Sarah Paisdzior
Institute of Experimental Pediatric Endocrinology
Charité – Universitätsmedizin Berlin,
Augustenburger Platz 1
13353 Berlin
Germany   
Phone: +49 30 450 559 828   

Publication History

Received: 30 September 2021
Received: 19 November 2021

Accepted: 25 November 2021

Article published online:
07 March 2022

© 2022. Thieme. All rights reserved.

Georg Thieme Verlag KG
Rüdigerstraße 14, 70469 Stuttgart, Germany


Zoom Image
Fig. 1 (a) Functional domains of TRα1 and TRα2 isoforms. Published disease-causing variants are marked on the protein domain structure. (b) TRα isoforms resulting from alternative splicing, or different transcription start points, leading to proteins of different molecular weight. (A/B: a regulatory domain; DBD: DNA-binding domain; LBD: ligand-binding domain)
Zoom Image
Fig. 2 (a) The DNA binding domains (DBD) of the unliganded TRα1 homodimer bind to responsive elements of the thyroid hormone receptor (TRE) in the promoter region of target genes. The ligand-binding domains (LBD) are in complex with co-repressors (CoRs) and target gene expression is inhibited. (b) Upon binding of thyroid hormone T3, CoRs are exchanged for co-activators (CoA), and the target genes are then expressed. (c) The weak antagonistic effect of TRα2 is probably due to competition for TREs between both isoforms without interaction with CoRs.