Bitter taste evolution

Last updated

The evolution of bitter taste receptors has been one of the most dynamic evolutionary adaptations to arise in multiple species. This phenomenon has been widely studied in the field of evolutionary biology because of its role in the identification of toxins often found on the leaves of inedible plants. A palate more sensitive to these bitter tastes would, theoretically, have an advantage over members of the population less sensitive to these poisonous substances because they would be much less likely to ingest toxic plants. Bitter-taste genes have been found in a host of vertebrates, including sharks and rays, [1] and the same genes have been well characterized in several common laboratory animals such as primates and mice, as well as in humans. The primary gene responsible for encoding this ability in humans is the TAS2R gene family which contains 25 functional loci as well as 11 pseudogenes. The development of this gene has been well characterized, with proof that the ability evolved before the human migration out of Africa. [2] The gene continues to evolve in the present day.

Contents

TAS2R

The bitter taste receptor family, T2R (TAS2R), is encoded on chromosome 7 and chromosome 12. Genes on the same chromosome have shown remarkable similarity with each other, suggesting that the primary mutagenic forces in evolution of TAS2R are duplication events. These events have occurred in at least seven primate species: chimpanzee, human, gorilla, orangutan, rhesus macaque and baboon. [3] The high variety among primate and rodent populations additionally suggests that, while selective constraint on these genes certainly exists, its effect is rather slight.

Members of the T2R family encode alpha subunits of G-protein-coupled receptors, which are involved in intracellular taste transduction, not only on the taste buds but also in the pancreas and gastrointestinal tract. The mechanism of transduction is shown by exposure of the endocrine and gastrointestinal cells containing the receptors to bitter compounds, most famously phenylthiocarbamide (PTC). Exposure to PTC causes an intracellular cascade as evidenced by a large and rapid increase in intracellular calcium ions. [4]

Toxins as the primary selective force

The primary selective adaptation that arises from bitter taste is to detect poisonous compounds, as most poisonous compounds in nature are bitter. However, this trait is not always advantageous, as bitter compounds exist in nature that are not poisonous. Exclusive rejection of these compounds would in fact be a disadvantageous trait, as it would make it more difficult to find food. Toxic and bitter compounds do, however, exist in different diets at different frequencies. [5] Sensitivities to bitter compounds should follow the requirements of different diets logically, as species that can afford to reject plants due to their low plant diet (carnivores) have a higher sensitivity to bitter compounds than those that exclusively ingest plants. Exposure to the bitter marker quinine hydrochloride supported this fact, as the sensitivities to bitter compounds were highest in carnivores, followed by omnivores, then grazers and browsers. [6] This identifies toxic plants as the primary selective force for bitter taste.

This phenomenon is confirmed with genetic analysis. One measure of positive selection is Ka/Ks, the ratio of synonymous to non-synonymous mutations. If the rate of synonymous mutation is higher than the rate of non-synonymous mutation, then the trait created by the non-synonymous mutation is being selected for relative to the neutral synonymous mutations. For the bitter taste gene family, TAS2R, this ratio is over one in the loci responsible for the extracellular binding domains of the receptors. [7] This indicates that the part of the receptor responsible for binding the bitter ligands is under positive selective pressure.

TAS2R development in human history

The pseudogenes mentioned earlier are produced by a number of gene silencing events, the rate of which is constant throughout primate species. Several of these pseudogenes maintain a role in modulating taste response, however. By studying the silencing events in humans, it is possible to theorize the selective pressures on humans throughout their evolutionary history. As is the case with the usual distribution of human genetic variation, the highest rate of diversity in TAS2R pseudogenes was often found in African populations. This was not the case with two pseudogene loci: TAS2R6P and TAS2R18P, where the highest diversity was found in non-African populations. This suggests that the functional versions of these genes arose before the human migration out of Africa into an area where selective constraint did not remove non-functional versions of these gene loci. This allowed the pseudogene frequency to increase, creating genetic variance at those loci. [2] This is an example of relaxed environmental constraint allowing silencing mutations to lead to pseudogenization of once important loci.

The gene locus, TAS2R16, also tells a story about bitter taste evolution. Varying rates of positive selection in different areas of the world give an indication of the selective pressures and events in those areas. At this locus, the 172Asn allele is the most common, especially in areas of Eurasia and in pygmy tribes in Africa, where it is nearly fixed. This suggests that the gene has had a relaxed selective constraint in most areas of Africa in comparison to Eurasia. This has been attributed to the increased knowledge of toxic plants in the area that arose around 10,000 years ago. The increased frequency of 172Asn in Eurasia suggests that the migration out of Africa into areas with different climates and foliage rendered the knowledge of toxic plants in Africa useless, forcing the populations to rely once again on the 172Asn allele, causing higher rates of positive selection. The high rate of 172Asn in Pygmy populations is more difficult to explain. The effective population size of these isolated populations is quite small, indicating that genetic drift explained by the founder effect is the cause of these atypically high rates. [8] The different environments that have contained humans have placed different levels of selection on the population, forcing a wide variety in at the TAS2R loci across humanity.

Relaxed constraint

Neutral evolution in the bitter taste trait in humans is well documented by evolutionary biologists. In all human populations there have been high rates of synonymous and non-synonymous substitutions that cause pseudogenization. These events cause alleles that are present to this day because of relaxed selective constraint by the environment. The genes under neutral evolution in humans are very similar to several genes in chimpanzees in both their synonymous and non-synonymous mutation rates, suggesting that relaxed selective constraint started before the divergence of the two species. [9]

The cause of this relaxed constraint was primarily in lifestyle changes in hominids. Roughly two million years ago, the hominid diet shifted from a primarily vegetarian diet to an increasingly meat-based diet. This led to a reduction in the amount of toxic foods regularly encountered by humanity's early ancestors. Additionally, the use of fire began around 800,000 years ago, which further detoxified food and led to a decreased dependence on TAS2R to detect poisonous food. Evolutionary biologists have theorized how, with fire being an exclusively human tool, relaxed selective constraint has been found in chimpanzees as well. Meat does account for about 15% of the chimpanzee diet, with much of the other 85% being made up of ripe fruits, which very rarely contains toxins. This comes in contrast to other primates whose diets are entirely composed of leaves, unripe fruits, and bark, which have comparatively high levels of toxins. [9] The differences in diets between chimpanzees and other primates accounts for the different levels of selective constraint.

Related Research Articles

<span class="mw-page-title-main">Human genome</span> Complete set of nucleic acid sequences for humans

The human genome is a complete set of nucleic acid sequences for humans, encoded as DNA within the 23 chromosome pairs in cell nuclei and in a small DNA molecule found within individual mitochondria. These are usually treated separately as the nuclear genome and the mitochondrial genome. Human genomes include both protein-coding DNA sequences and various types of DNA that does not encode proteins. The latter is a diverse category that includes DNA coding for non-translated RNA, such as that for ribosomal RNA, transfer RNA, ribozymes, small nuclear RNAs, and several types of regulatory RNAs. It also includes promoters and their associated gene-regulatory elements, DNA playing structural and replicatory roles, such as scaffolding regions, telomeres, centromeres, and origins of replication, plus large numbers of transposable elements, inserted viral DNA, non-functional pseudogenes and simple, highly repetitive sequences. Introns make up a large percentage of non-coding DNA. Some of this non-coding DNA is non-functional junk DNA, such as pseudogenes, but there is no firm consensus on the total amount of junk DNA.

<span class="mw-page-title-main">Neutral theory of molecular evolution</span> Theory of evolution by changes at the molecular level

The neutral theory of molecular evolution holds that most evolutionary changes occur at the molecular level, and most of the variation within and between species are due to random genetic drift of mutant alleles that are selectively neutral. The theory applies only for evolution at the molecular level, and is compatible with phenotypic evolution being shaped by natural selection as postulated by Charles Darwin.

<span class="mw-page-title-main">Pseudogene</span> Functionless relative of a gene

Pseudogenes are nonfunctional segments of DNA that resemble functional genes. Most arise as superfluous copies of functional genes, either directly by gene duplication or indirectly by reverse transcription of an mRNA transcript. Pseudogenes are usually identified when genome sequence analysis finds gene-like sequences that lack regulatory sequences needed for transcription or translation, or whose coding sequences are obviously defective due to frameshifts or premature stop codons. Pseudogenes are a type of junk DNA.

Phenylthiocarbamide (PTC), also known as phenylthiourea (PTU), is an organosulfur thiourea containing a phenyl ring.

<i>The Neutral Theory of Molecular Evolution</i>

The Neutral Theory of Molecular Evolution is an influential monograph written in 1983 by Japanese evolutionary biologist Motoo Kimura. While the neutral theory of molecular evolution existed since his article in 1968, Kimura felt the need to write a monograph with up-to-date information and evidences showing the importance of his theory in evolution.

Human evolutionary genetics studies how one human genome differs from another human genome, the evolutionary past that gave rise to the human genome, and its current effects. Differences between genomes have anthropological, medical, historical and forensic implications and applications. Genetic data can provide important insights into human evolution.

<span class="mw-page-title-main">Taste receptor</span> Type of cellular receptor that facilitates taste

A taste receptor or tastant is a type of cellular receptor which facilitates the sensation of taste. When food or other substances enter the mouth, molecules interact with saliva and are bound to taste receptors in the oral cavity and other locations. Molecules which give a sensation of taste are considered "sapid".

<span class="mw-page-title-main">TAS2R38</span> Protein-coding gene in the species Homo sapiens

Taste receptor 2 member 38 is a protein that in humans is encoded by the TAS2R38 gene. TAS2R38 is a bitter taste receptor; varying genotypes of TAS2R38 influence the ability to taste both 6-n-propylthiouracil (PROP) and phenylthiocarbamide (PTC). Though it has often been proposed that varying taste receptor genotypes could influence tasting ability, TAS2R38 is one of the few taste receptors shown to have this function.

<span class="mw-page-title-main">TAS2R10</span> Protein-coding gene in the species Homo sapiens

Taste receptor type 2 member 10 is a protein that in humans is encoded by the TAS2R10 gene. The protein is responsible for bitter taste recognition in mammals. It serves as a defense mechanism to prevent consumption of toxic substances which often have a characteristic bitter taste.

<span class="mw-page-title-main">TAS2R14</span> Protein-coding gene in the species Homo sapiens

Taste receptor type 2 member 14 is a protein that in humans is encoded by the TAS2R14 gene.

<span class="mw-page-title-main">TAS2R41</span> Protein-coding gene in the species Homo sapiens

Taste receptor type 2 member 41 is a protein that in humans is encoded by the TAS2R41 gene.

<span class="mw-page-title-main">TAS2R43</span> Protein-coding gene in the species Homo sapiens

Taste receptor type 2 member 43 is a protein that in humans is encoded by the TAS2R43 gene.

<span class="mw-page-title-main">TAS2R31</span> Protein-coding gene in the species Homo sapiens

Taste receptor, type 2, member 31, also known as TAS2R31, is a protein which in humans is encoded by the TAS2R31 gene. This bitter taste receptor has been shown to respond to saccharin in vitro.

<span class="mw-page-title-main">TAS2R46</span> Protein-coding gene in the species Homo sapiens

Taste receptors for bitter substances (T2Rs/TAS2Rs) belong to the family of G-protein coupled receptors and are related to class A-like GPCRs. There are 25 known T2Rs in humans responsible for bitter taste perception.

<span class="mw-page-title-main">TAS2R20</span> Protein-coding gene in the species Homo sapiens

Taste receptor type 2 member 20 is a protein that in humans is encoded by the TAS2R20 gene.

<span class="mw-page-title-main">Evolution of color vision in primates</span> Loss and regain of colour vision during the evolution of primates

The evolution of color vision in primates is highly unusual compared to most eutherian mammals. A remote vertebrate ancestor of primates possessed tetrachromacy, but nocturnal, warm-blooded, mammalian ancestors lost two of four cones in the retina at the time of dinosaurs. Most teleost fish, reptiles and birds are therefore tetrachromatic while most mammals are strictly dichromats, the exceptions being some primates and marsupials, who are trichromats, and many marine mammals, who are monochromats.

A nonsynonymous substitution is a nucleotide mutation that alters the amino acid sequence of a protein. Nonsynonymous substitutions differ from synonymous substitutions, which do not alter amino acid sequences and are (sometimes) silent mutations. As nonsynonymous substitutions result in a biological change in the organism, they are subject to natural selection.

A conserved non-coding sequence (CNS) is a DNA sequence of noncoding DNA that is evolutionarily conserved. These sequences are of interest for their potential to regulate gene production.

Odor molecules are detected by the olfactory receptors in the olfactory epithelium of the nasal cavity. Each receptor type is expressed within a subset of neurons, from which they directly connect to the olfactory bulb in the brain. Olfaction is essential for survival in most vertebrates; however, the degree to which an animal depends on smell is highly varied. Great variation exists in the number of OR genes among vertebrate species, as shown through bioinformatic analyses. This diversity exists by virtue of the wide-ranging environments that they inhabit. For instance, dolphins that are secondarily adapted to an aquatic niche possess a considerably smaller subset of genes than most mammals. OR gene repertoires have also evolved in relation to other senses, as higher primates with well-developed vision systems tend to have a smaller number of OR genes. As such, investigating the evolutionary changes of OR genes can provide useful information on how genomes respond to environmental changes. Differences in smell sensitivity are also dependent on the anatomy of the olfactory apparatus, such as the size of the olfactory bulb and epithelium.

<span class="mw-page-title-main">PTC tasting</span>

PTC tasting is a classic genetic marker in human population genetics investigations.

References

  1. Itoigawa, Akihiro; Toda, Yasuka; Kuraku, Shigehiro; Ishimaru, Yoshiro (2024-04-08). "Evolutionary origins of bitter taste receptors in jawed vertebrates". Current Biology. 34 (7): R271–R272. doi: 10.1016/j.cub.2024.02.024 .
  2. 1 2 Davide Risso; Sergio Tofanelli; Gabriella Morini; Donata Luiselli & Dennis Drayna (2014). "Genetic variation in taste receptor pseudogenes provides evidence for a dynamic role in human evolution". BMC Evolutionary Biology . 14 (1): 198. Bibcode:2014BMCEE..14..198R. doi: 10.1186/s12862-014-0198-8 . PMC   4172856 . PMID   25216916.
  3. Anne Fischer; Yoav Gilad; Orna Man & Svante Pääbo (2004). "Evolution of bitter taste receptors in humans and apes". Molecular Biology and Evolution . 22 (3): 432–436. doi: 10.1093/molbev/msi027 . PMID   15496549.
  4. S. Vincent Wu; Nora Rozengurt; Moon Yang; Steven H. Young; James Sinnett-Smith & Enrique Rozengurt (2001). "Expression of bitter taste receptors of the T2R family in the gastrointestinal tract and enterendocrine STC-1 cells". Proceedings of the National Academy of Sciences of the United States of America . 99 (4): 2392–2397. doi: 10.1073/pnas.042617699 . PMC   122375 . PMID   11854532.
  5. Sambu, Sammy (3 December 2019). "The determinants of chemoreception as evidenced by gradient boosting machines in broad molecular fingerprint spaces". PeerJ Organic Chemistry. 1: e2. doi: 10.7717/peerj-ochem.2 .
  6. John I. Glendinning (1994). "Is the bitter rejection response always adaptive?". Physiology & Behavior . 56 (6): 1217–1222. doi:10.1016/0031-9384(94)90369-7. PMID   7878094. S2CID   22945002.
  7. Peng Shi; Jianzhi Zhang; Hui Yang & Ya-ping Zhang (2003). "Adaptive diversification of bitter taste receptor genes in mammalian evolution". Molecular Biology and Evolution . 20 (5): 805–814. doi: 10.1093/molbev/msg083 . PMID   12679530.
  8. Hui Li; Andrew J. Pakstis; Judith R. Kidd & Kenneth K. Kidd (2011). "Selection on the human bitter taste gene, TAS2R16, in Eurasian populations". Human Biology . 83 (3): 363–377. doi:10.3378/027.083.0303. PMID   21740153. S2CID   15490534.
  9. 1 2 Xiaoxia Wang; Stephanie D. Thomas & Jianzhi Zhang (2004). "Relaxation of selective constraint and loss of function in the evolution of human bitter taste receptor genes". Human Molecular Genetics . 13 (21): 2671–2678. doi: 10.1093/hmg/ddh289 . PMID   15367488.