Chapter Nine - The SLC16A Family of Monocarboxylate Transporters (MCTs)—Physiology and Function in Cellular Metabolism, pH Homeostasis, and Fluid Transport

https://doi.org/10.1016/B978-0-12-394316-3.00009-0Get rights and content

Abstract

The SLC16A family of monocarboxylate transporters (MCTs) is composed of 14 members. MCT1 through MCT4 (MCTs 1–4) are H+-coupled monocarboxylate transporters, MCT8 and MCT10 transport thyroid hormone and aromatic amino acids, while the substrate specificity and function of other MCTs have yet to be determined. The focus of this review is on MCTs 1–4 because their role in lactate transport is intrinsically linked to cellular metabolism in various biological systems, including skeletal muscle, brain, retina, and testis. Although MCTs 1–4 all transport lactate, they differ in their transport kinetics and vary in tissue and subcellular distribution, where they facilitate “lactate-shuttling” between glycolytic and oxidative cells within tissues and across blood–tissue barriers. However, the role of MCTs 1–4 is not confined to cellular metabolism. By interacting with bicarbonate transport proteins and carbonic anhydrases, MCTs participate in the regulation of pH homeostasis and fluid transport in renal proximal tubule and corneal endothelium, respectively. Here, we provide a comprehensive review of MCTs 1–4, linking their cellular distribution to their functions in various parts of the human body, so that we can better understand the physiological roles of MCTs at the systemic level.

Introduction

The SLC16 family of monocarboxylate transporters, MCTs, is composed of 14 members. They share similarities in sequence and protein structure, but each member differs from others in the substrates that they transport. Lactate, among other monocarboxylates, is the primary substrate for MCTs 1–4. MCT6 was shown to transport synthetic drugs such as probenecid and bumetanide. MCT8 transports thyroid hormones (T3 and T4) and mutations in this transporter (Allan–Herndon–Dudley Syndrome) cause crippling defects in brain and motor functions throughout development because of the lack of T3 uptake by neurons (Friesema et al., 2004; Schwartz et al., 2005). MCT10 (or T-type amino acid transporter-1; TAT1) is also capable of transporting thyroid hormones (Friesema et al., 2008), but unlike MCT8, its transport substrate includes other aromatic amino acids and that its gene expression is found primarily in the kidney, skeletal muscle, and small intestine (Kim et al., 2002). MCT9 was recently identified as a carnitine transporter through a meta-analysis of 14 GWAS studies of patients with elevated serum uric acid (Kolz et al., 2009). This finding was also independently verified in a separate study that evaluated the genetic links to human metabolism by coupling GWAS analysis with a high-throughput screening of blood metabolites (Suhre et al., 2011). Beyond the abovementioned MCTs, however, the transport substrates of the other MCTs (MCTs 5, 7, 11–14) are unknown. Of all MCTs, MCTs 1–4 are the most widely studied and understood. This is not surprising, as lactate has well-established roles in acid–base balance, with the earliest studies dating back to almost half a century ago. The advent of a noninvasive fluorescence-based intracellular pH indicator, BCECF (Rink, Tsien, & Pozzan, 1982), allowed physiologists to easily demonstrate that transmembrane lactate transport is accompanied by intracellular acidification. This observation gave rise to an explosion of studies that demonstrated the existence and function of H+-coupled lactate transporters in various biological systems (Giasson & Bonanno, 1994; Kenyon, Yu, La Cour, & Miller, 1994; Poole & Halestrap, 1991; Roth & Brooks, 1990; Wang, Levi, & Halestrap, 1994).

For many decades, lactate was regarded as nothing more than a metabolic waste product of glycolysis. The accumulation of lactate within cells and their microenvironment, such as in muscle fibers during exercise, was thought to cause lactic acidosis with detrimental consequences to normal physiological function. However, with accumulating evidence demonstrating that lactate can be utilized as a metabolic substrate for oxidative respiration, there came a realization that lactate is more than simply a waste product (Gladden, 2004)—a radical idea at the time that raised much interest in the roles of MCTs 1–4 in metabolism in many parts of the human body, including the muscle, brain, kidney, liver, retina, and testis, all of which will be discussed in detail below (Section 2). Although MCTs are important for normal metabolic function, they are also implicated in disease. In cancer cells, for example, MCT1 and MCT4 are commonly upregulated to adapt to a change in metabolism and to maintain pH homeostasis. In addition to regulating cellular metabolism, MCTs also participate in fluid transport in corneal endothelium of the eye (Section 4.2) and acid–base balance in the proximal convoluted tubule (PCT) (Section 4.3) of the kidney. These processes are achieved by functional interaction between MCTs and HCO3 transport mechanisms (Section 3). With most of the functions of MCTs 1–4 revolving around lactate, it is easy to overlook the fact that MCTs 1–4 are just as capable of transporting other substrates, such as acetate, propionate, and butyrate—all of which are absorbed from the colon lumen via Na+-coupled MCTs (SMCTs) and MCT1 (Section 4.4).

Although the crystal structure of MCTs has not been determined, hydrophobicity plots of MCTs 1–14 predict that these proteins consist of 12 transmembrane domains (TMs) with cytoplasmic N and C termini (Halestrap, 2012). The coupling of H+ to monocarboxylate transport is unique to MCTs 1–4, and the structural property that is responsible for this feature, a lysine residue in TM1 (Lys38; Fig. 1A), has only been recently identified in rat MCT1 (Wilson, Meredith, Bunnun, Sessions, & Halestrap, 2009). In their model, Lys38 is normally uncharged in the hydrophobic environment of the ion pore, but it can bind to an extracellular H+ to trigger the open conformation, in which the same lysine residue could subsequently bind a lactate ion. Upon binding, both the H+ and lactate ions are transferred to an aspartate–arginine ion pair (Asp302–Arg306; TM8) located deeper in the channel pore (Manoharan, Wilson, Sessions, & Halestrap, 2006). These residues were identified in rat MCT1, but they are also found in MCTs 1–4 of both rat and human (Fig. 1B). Although MCTs 1–4 transport lactate and other similar substrates, they differ in substrate binding affinity and selectivity. An early study by Garcia et al. identified an amino acid residue that regulates the substrate selectivity of MCT1—a phenylalanine residue in TM10, Phe360 (Garcia, Brown, Pathak, & Goldstein, 1995). It is interesting to note that in MCTs 3 and 4, however, a tyrosine, rather than phenylalanine, is found in this position (Fig. 1C). This switch may account for the differences in lactate-binding affinities among MCTs 1–4, which seem to correlate with their function when expressed in specific cell types within a biological system. For example, MCT1 and MCT2, with their high affinity for lactate (Km ≈ 1–3 mM), possess the ability to mediate rapid lactate uptake into cells (Halestrap & Price, 1999). As such, MCT1 and MCT2 are commonly expressed in tissues that utilize lactate as a substrate for oxidative metabolism, such as slow-twitch type I muscle fibers (MCT1) and neurons (MCT2). On the other hand, MCT3 and MCT4 have lower affinities for lactate (MCT3, Km ≈ 6 mM; MCT4, Km ≈ 30 mM), but they possess a wide range of transport rates and are, therefore, useful for maintaining lactate equilibrium in cells. MCT4 is commonly expressed in glycolytic cell types, such as type II fast-twitch skeletal muscle cells and astrocytes, where it mediates lactic acid extrusion. MCT3 is preferentially expressed in the retinal pigment epithelium, where it also mediates lactic acid extrusion.

With specific MCTs being differentially expressed in different cell types, or even in different membranes of a polarized epithelium, it is obvious that MCTs, similar to many other transporters, are under strict regulatory control. SLC16A1 (MCT1) and SLC16A3 (MCT4) carry multiple splice variants and differ only at the 5′-UTR region, suggesting that MCT1 and MCT4 transcripts can be subject to both transcriptional and post-transcriptional regulation. This level of control may be necessary because MCT1 and MCT4 are widely expressed in many different cell types, thus requiring additional mechanisms to regulate their relative abundance in a cell-specific manner. Furthermore, MCT1 transcripts can also be regulated by microRNAs, which target the mRNA at its 3′-UTR for translational repression or mRNA degradation. This mechanism is particularly important in beta-cells, where MCT1 is not normally expressed (Pullen, da Silva Xavier, Kelsey, & Rutter, 2011). On the other hand, no other transcript variants for SLC16A7 (MCT2) and SLC16A8 (MCT3) have been found, consistent with the more limited tissue distributions of MCT2 and MCT3.

Although the protein abundance of MCTs 1–4 can be regulated at the genetic level, the proper folding and trafficking of the MCTs 1–4 proteins to the plasma membrane requires each MCT to be assembled with either CD147/basigin (encoded by Bsg) or embigin (encoded by Emb), both of which are highly glycosylated single-pass membrane proteins that serve as chaperones for proper targeting of MCTs to the plasma membrane (Kirk et al., 2000; Philp, Ochrietor, Rudoy, Muramatsu, & Linser, 2003; Wilson et al., 2005). CD147 is widely expressed in tissues and forms a heterodimer with MCT1, 3, or 4. Embigin, on the other hand, associates primarily with MCT2, but it can also interact with MCT1 in some tissues. The dependence of MCTs 1–4 on their accessory proteins has been demonstrated in in vitro studies, but the physiological importance of this interaction in MCT maturation and trafficking is perhaps best demonstrated in the Bsg-null mouse (Nakai, Chen, & Nowak, 2006; Philp et al., 2003), which present with blindness, sterility, immunodeficiency, and problems with learning and memory (Hori et al., 2000; Igakura et al., 1998; Naruhashi et al., 1997). These phenotypes, as is discussed in the following sections (Sections 2 and 4), are direct pathological manifestations caused by the absence of MCTs. The underlying mechanisms behind these conditions will become readily apparent when these MCTs are placed in the context of their physiological functions in various cell types within their native biological systems.

Section snippets

Skeletal Muscle

The concept of lactate-shuttling is perhaps best established in skeletal muscle, which is composed of two major cell types with vastly different metabolic properties: type I oxidative fibers and type II glycolytic fibers (reviewed in (Schiaffino & Reggiani, 2011)). Anatomically, type I and II muscle fibers are closely apposed onto each other and separated by a small volume of extracellular space that is occupied by neurovascular structures. Type I oxidative fibers can be distinguished from type

Interactions between MCTs and Bicarbonate Transport Mechanisms

MCT-mediated transfer of lactate between cells and their microenvironment often involves large acid loads that can inadvertently disrupt cellular function. To avoid such an event, most biological systems employ a set of pH-regulatory systems to help buffer transient or chronic changes in intracellular pH. As in other mammals, CO2/HCO3 buffering is the major pH-regulatory system in humans, and it is mediated by the activities of carbonic anhydrases (CAs) and HCO3 transporters (reviewed in (

Retinal Pigment Epithelium: pH Homeostasis and Metabolic Acid Removal

As previously discussed (Section 2.4), the neural retina is a metabolically robust tissue that generates large quantities of both CO2/HCO3 and lactic acid, both of which are released into the photoreceptor extracellular milieu (SRS) where they are subsequently removed by the RPE and transported into the choroidal blood supply. Earlier studies demonstrated that CO2 and lactate levels in the SRS could reach up to 10% and 13 mM, respectively, both of which are significantly higher than that in

Future Directions

Studies in the muscle and brain have established MCTs 1–4 as integral components of the lactate-shuttle network that are directly linked to cellular metabolism. With the demonstration that MCTs can form a functional complex with HCO3 transporters and carbonic anhydrases, it becomes clear that MCTs also participate in pH regulation and homeostasis, as well as transepithelial fluid transport in a variety of tissues, such as the kidney, cornea, and retina/RPE. Hence, it is not surprising that the

Conclusion

Since Poole and Halestrap published their first paper on the isolation of the erythrocyte lactate transporter (Poole & Halestrap, 1988), which was subsequently found to be identical to MCT1 cloned by Garcia, Goldstein, Pathak, Anderson, and Brown (1994), there has been renewed interest in these transporters and their function in metabolism in both health and disease. The cloning of MCT1 (SCL16A1) paved the way for identification of other members of the MCT family and the development of

Acknowledgments

The authors thank Melissa Liu and Lea Hecht for reading the manuscript and their insightful comments.

This work was funded by NIH grant EY-012042.

References (169)

  • G.M. Feldman

    HCO3 secretion by rat distal colon: effects of inhibitors and extracellular Na+

    Gastroenterology

    (1994)
  • E.C. Friesema et al.

    Association between mutations in a thyroid hormone transporter and severe X-linked psychomotor retardation

    Lancet

    (2004)
  • M. Galie et al.

    Epithelial and mesenchymal tumor compartments exhibit in vivo complementary patterns of vascular perfusion and glucose metabolism

    Neoplasia

    (2007)
  • C.K. Garcia et al.

    cDNA cloning of MCT2, a second monocarboxylate transporter expressed in different cells than MCT1

    Journal of Biological Chemistry

    (1995)
  • C.K. Garcia et al.

    Molecular characterization of a membrane transporter for lactate, pyruvate, and other monocarboxylates: implications for the Cori cycle

    Cell

    (1994)
  • L.R. Gawenis et al.

    Colonic anion secretory defects and metabolic acidosis in mice lacking the NBC1 Na+/HCO3 cotransporter

    Journal of Biological Chemistry

    (2007)
  • C. Giasson et al.

    Facilitated transport of lactate by rabbit corneal endothelium

    Experimental Eye Research

    (1994)
  • E. Gopal et al.

    Cloning and functional characterization of human SMCT2 (SLC5A12) and expression pattern of the transporter in kidney

    Biochimica et Biophysica Acta

    (2007)
  • J.D. Gordan et al.

    HIF and c-Myc: sibling rivals for control of cancer cell metabolism and proliferation

    Cancer Cell

    (2007)
  • S. Hamann et al.

    Cotransport of H+, lactate, and H2O in porcine retinal pigment epithelial cells

    Experimental Eye Research

    (2003)
  • D. Hanahan et al.

    Hallmarks of cancer: the next generation

    Cell

    (2011)
  • T. Igakura et al.

    A null mutation in basigin, an immunoglobulin superfamily member, indicates its important roles in peri-implantation development and spermatogenesis

    Developmental Biology

    (1998)
  • D.K. Kim et al.

    The human T-type amino acid transporter-1: characterization, gene organization, and chromosomal location

    Genomics

    (2002)
  • M. Klier et al.

    Transport activity of the high-affinity monocarboxylate transporter MCT2 is enhanced by extracellular carbonic anhydrase IV but not by intracellular carbonic anhydrase II

    Journal of Biological Chemistry

    (2011)
  • J. Li et al.

    Dependence of cAMP meditated increases in Cl and HCO(3)− permeability on CFTR in bovine corneal endothelial cells

    Experimental Eye Research

    (2008)
  • G. Lonnerholm et al.

    Amount and distribution of carbonic anhydrases CA I and CA II in the gastrointestinal tract

    Gastroenterology

    (1985)
  • C.V. Lowry et al.

    Enzyme patterns in single human muscle fibers

    Journal of Biological Chemistry

    (1978)
  • Y.F. Lo et al.

    Severe metabolic acidosis causes early lethality in NBC1 W516X knock-in mice as a model of human isolated proximal renal tubular acidosis

    Kidney International

    (2011)
  • J. Lu et al.

    Effect of human carbonic anhydrase II on the activity of the human electrogenic Na/HCO3 cotransporter NBCe1-A in Xenopus oocytes

    Journal of Biological Chemistry

    (2006)
  • J.E. Melvin et al.

    Mouse down-regulated in adenoma (DRA) is an intestinal Cl(−)/HCO(3)(−) exchanger and is up-regulated in colon of mice lacking the NHE3 Na(+)/H(+) exchanger

    Journal of Biological Chemistry

    (1999)
  • J. Adijanto et al.

    CO2-induced ion and fluid transport in human retinal pigment epithelium

    Journal of General Physiology

    (2009)
  • A.J. Adler et al.

    Distribution of glucose and lactate in the interphotoreceptor matrix

    Ophthalmic Research

    (1992)
  • B.V. Alvarez et al.

    Direct extracellular interaction between carbonic anhydrase IV and the human NBC1 sodium/bicarbonate co-transporter

    Biochemistry

    (2003)
  • D. Attwell et al.

    Neuroenergetics and the kinetic design of excitatory synapses

    Nature Reviews Neuroscience

    (2005)
  • H.M. Becker et al.

    Intramolecular proton shuttle supports not only catalytic but also noncatalytic function of carbonic anhydrase II

    Proceedings of the National Academy of Sciences of the United States of America

    (2011)
  • H.M. Becker et al.

    Localization of members of MCT monocarboxylate transporter family Slc16 in the kidney and regulation during metabolic acidosis

    American Journal of Physiology. Renal Physiology

    (2010)
  • R. Bellomo

    Bench-to-bedside review: lactate and the kidney

    Critical Care

    (2002)
  • L. Bergersen et al.

    A novel postsynaptic density protein: the monocarboxylate transporter MCT2 is co-localized with delta-glutamate receptors in postsynaptic densities of parallel fiber-Purkinje cell synapses

    Experimental Brain Research

    (2001)
  • R. Bernards et al.

    A progression puzzle

    Nature

    (2002)
  • C.X. Bittner et al.

    High resolution measurement of the glycolytic rate

    Front Neuroenergetics

    (2010)
  • R. Boidot et al.

    Regulation of monocarboxylate transporter MCT1 expression by p.53 mediates inward and outward lactate fluxes in tumors

    Cancer Research

    (2012)
  • W.F. Boron

    Acid-base transport by the renal proximal tubule

    Journal of the American Society of Nephrology

    (2006)
  • W.F. Boron et al.

    Modular structure of sodium-coupled bicarbonate transporters

    Journal of Experimental Biology

    (2009)
  • J.R. Casey

    Why bicarbonate?

    Biochemistry and Cell Biology

    (2006)
  • J.R. Casey et al.

    Bicarbonate homeostasis in excitable tissues: role of AE3 Cl-/HCO3- exchanger and carbonic anhydrase XIV interaction

    American Journal of Physiology. Cell Physiology

    (2009)
  • A.N. Charney et al.

    Nonionic diffusion of short-chain fatty acids across rat colon

    American Journal of Physiology

    (1998)
  • F. Chemello et al.

    Microgenomic analysis in skeletal muscle: expression signatures of individual fast and slow myofibers

    PLoS One

    (2011)
  • C.Y. Cheng et al.

    An intracellular trafficking pathway in the seminiferous epithelium regulating spermatogenesis: a biochemical and molecular perspective

    Critical Reviews In Biochemistry and Molecular Biology

    (2009)
  • J.L. Chen et al.

    The genomic analysis of lactic acidosis and acidosis response in human cancers

    PLoS Genetics

    (2008)
  • G. Chidlow et al.

    Expression of monocarboxylate transporters in rat ocular tissues

    American Journal of Physiology. Cell Physiology

    (2005)
  • Cited by (78)

    • The Warburg-like effect in male reproductive events

      2023, Glycolysis: Tissue-Specific Metabolic Regulation in Physio-pathological Conditions
    • Aging astrocytes metabolically support aging axon function by proficiently regulating astrocyte-neuron lactate shuttle

      2022, Experimental Neurology
      Citation Excerpt :

      Endothelial cells as part of the BBB express MCT1 in luminal and abluminal surfaces (Bergersen, 2015; Pellerin et al., 1998). Under physiological conditions, astrocytes produce and release lactate in the extracellular matrix but in order to maintain low extracellular levels, excess lactate is transferred into the blood circulation through MCT-1 expressed at the surface of endothelial cells (Adijanto and Philp, 2012). The impact of aging on MCT1 expression, endothelial regulation of lactate, and oligodendrocyte-axon lactate shuttle efficiency are yet to be explored.

    • Embigin facilitates monocarboxylate transporter 1 localization to the plasma membrane and transition to a decoupling state

      2022, Cell Reports
      Citation Excerpt :

      In human, the rapid exchange of intracellular, metabolically important monocarboxylates, such as lactate, pyruvate, and ketone bodies, with the circulation is primarily mediated by proton-coupled monocarboxylate transporters (MCTs) encoded by SLC16 family members (Adijanto and Philp, 2012; Felmlee et al., 2020; Halestrap 2013a; Halestrap, 2013b).

    View all citing articles on Scopus
    View full text