Hostname: page-component-8448b6f56d-42gr6 Total loading time: 0 Render date: 2024-04-20T02:29:09.838Z Has data issue: false hasContentIssue false

β-Carotene in the human body: metabolic bioactivation pathways – from digestion to tissue distribution and excretion

Published online by Cambridge University Press:  12 February 2019

Torsten Bohn
Affiliation:
Luxembourg Institute of Health, rue 1 A-B Thomas Edison, L-1445 Strassen, Luxembourg
Charles Desmarchelier
Affiliation:
C2VN, Aix-Marseille Univ., INRA, INSERM, Marseille, France
Sedef N. El
Affiliation:
Engineering Faculty, Food Engineering Department, Ege University, Izmir, Turkey
Jaap Keijer
Affiliation:
Human and Animal Physiology, Wageningen University, Wageningen, The Netherlands
Evert van Schothorst
Affiliation:
Human and Animal Physiology, Wageningen University, Wageningen, The Netherlands
Ralph Rühl*
Affiliation:
Paprika Bioanalytics BT, Debrecen, Hungary
Patrick Borel
Affiliation:
C2VN, Aix-Marseille Univ., INRA, INSERM, Marseille, France
*
*Corresponding author: Ralph Rühl, email ralphruehl@web.de
Rights & Permissions [Opens in a new window]

Abstract

β-Carotene intake and tissue/blood concentrations have been associated with reduced incidence of several chronic diseases. Further bioactive carotenoid-metabolites can modulate the expression of specific genes mainly via the nuclear hormone receptors: retinoic acid receptor- and retinoid X receptor-mediated signalling. To better understand the metabolic conversion of β-carotene, inter-individual differences regarding β-carotene bioavailability and bioactivity are key steps that determine its further metabolism and bioactivation and mediated signalling. Major carotenoid metabolites, the retinoids, can be stored as esters or further oxidised and excreted via phase 2 metabolism pathways. In this review, we aim to highlight the major critical control points that determine the fate of β-carotene in the human body, with a special emphasis on β-carotene oxygenase 1. The hypothesis that higher dietary β-carotene intake and serum level results in higher β-carotene-mediated signalling is partly questioned. Alternative autoregulatory mechanisms in β-carotene / retinoid-mediated signalling are highlighted to better predict and optimise nutritional strategies involving β-carotene-related health beneficial mediated effects.

Type
Conference on ‘Nutrient–nutrient interaction’
Copyright
Copyright © The Authors 2019 

Introduction

β-Carotene is a tetra-terpenoid consisting of a C40 structure including two β-ionone rings. Together with lycopene, it is among the most frequently consumed dietary carotenoids in human subjects(Reference Biehler, Alkerwi and Hoffmann1Reference Wawrzyniak, Hamulka and Friberg3), also ranking among the highest in blood / plasma concentrations(Reference Wawrzyniak, Hamulka and Friberg3, Reference Fraser, Jaceldo-Siegl and Henning4). β-Carotene is the most important source of non-preformed vitamin A, as this molecule can, following absorption, be cleaved to form vitamin A (retinal). The maintenance of normal vision, enhancement of growth, tissue differentiation and reproduction was associated with β-carotene's dietary intake via fruits and vegetables as well as its blood concentrations, and have been further also associated with reduced incidence or disease biomarkers of several chronic complications such as type-2 diabetes(Reference Sluijs, Cadier and Beulens5, Reference Sugiura, Nakamura and Ogawa6) and other CVD(Reference Wang, Chung and McCullough7, Reference Karppi, Kurl and Ronkainen8).

However, highly dosed supplemental use of β-carotene has been correlated with smokers with negative outcomes, increasing total mortality(Reference Bjelakovic, Nikolova and Gluud9, Reference Bjelakovic, Nikolova and Gluud10) and enhancing lung cancer rate(Reference Omenn, Goodman and Thornquist11, 12). While nutritional relevant levels of β-carotene have been in the range of 0·1–8·8 (median 3·9) mg/d as reviewed recently(Reference Böhm, Borel and Corte-Real13), isolated administered β-carotene at rather high doses (>30 mg/d) may initiate previously mentioned negative effects. The reasons for these negative effects are not entirely clear but indicate that such high levels of β-carotene, which are usually non-harmful, may affect precancerous lesions, which are a hallmark of cancer development, and result from smoking(Reference van Helden, Godschalk and Swarts14). Studies in ferrets indicated altered non-beneficial retinoid signalling in the lung(Reference Wang, Liu and Bronson15). These studies suggest prudency with regards to form and dosing of β-carotene, while smoking appears as the major trigger of disease formation.

While earlier studies have emphasised direct antioxidant effects of β-carotene, including quenching singlet oxygen and lipid peroxides(Reference van Helden, Godschalk and Swarts14, Reference Krinsky and Johnson16, Reference Krinsky and Yeum17), more recent studies have especially highlighted that besides carotenoids, mainly carotenoid-metabolites obtain a role in altering gene expression(Reference Piga, van Dartel and Bunschoten18) reviewed in Kaulmann and Bohn(Reference Kaulmann and Bohn19), mainly via the interaction of the β-carotene- / retinol-metabolite all-trans-retinoic acid (ATRA)(Reference Rühl, Bub and Watzl20) with nuclear receptors, including retinoic acid receptor (RAR) and retinoid X receptor (RXR)(Reference Prakash, Liu and Hu21, Reference Ben-Dor, Nahum and Danilenko22). Nuclear hormone-mediated signalling is not directly but indirectly related to β-carotene intake. Initially, further cleavage and cleavage metabolites, especially the active vitamin A compounds (ATRA) are formed by β-carotene oxygenase 1 (BCO1) centric cleave or following eccentric cleavage by BCO2(Reference Amengual, Widjaja-Adhi and Rodriguez-Santiago23), which underlie a complex regulation in the organism, as explained further in the present paper. In addition and with an unclear nutritional and physiological relevance, transcription factors such as NF-κB and nuclear factor erythroid 2-related factor 2 (Nrf2) may also be involved in β-carotene-mediated signalling.

Recent studies, including gene-association studies(Reference Borel, Desmarchelier and Nowicki24Reference Borel, Desmarchelier and Nowicki26), have suggested that the bioavailability and bioactivity of β-carotene are related to key steps determining β-carotene absorption, distribution, metabolism and excretion (ADME), and are summarised as following:

  1. (a) the dietary release of β-carotene from the food matrix and its micellisation;

  2. (b) its cellular uptake into the enterocytes;

  3. (c) transport and metabolism of β-carotene, i.e.

    1. (c1) the intracellular metabolism of β-carotene in the enterocyte;

    2. (c2) or alternatively, the further transport of β-carotene within the organism and metabolism in target tissues;

  4. (d) further transport and bio-distribution of carotenoids or carotenoid-metabolites;

  5. (e) transmission and regulation of biological-mediated functions;

  6. (f) excretion.

Consequently, there is a large variability of β-carotene bioavailability in human subjects, due to both nutritional- and host-related factors, including genetic variations(Reference Desmarchelier and Borel27, Reference Bohn, Desmarchelier and Dragsted28).

In the present review and position paper, we aim to pinpoint and highlight these critical steps in the metabolism of β-carotene, which likely constitute the strongest levers regarding bioavailability and bioactivity. Focusing on gaps of knowledge, especially the further metabolism following cleavage by BCO1 and transmission of biological-mediated signalling and regulation/autoregulation of these pathways, are in the focus of this review.

β-Carotene during digestion

The majority of β-carotene in native plant matrices is present in the all-trans form(Reference Lessin and Schwartz29). Following food processing and especially heat treatment, various cis-isomers are formed(Reference Marx, Stuparic and Schieber30), such as the 9-cis, 13-cis and 15-cis isomers(Reference Lessin and Schwartz29), while other isomers are much less abundant. Processes such as novel emerging non-thermal food processing technologies, such as high-pressure processing, high-intensity pulsed electric fields and ultrasonication(Reference Cilla, Bosch and Barberá31) can cause some structural changes in carotenoids such as transcis isomerisation, potentially altering their solubility and their micellisation efficacy.

Following ingestion and mastication of the food matrix in the oral phase of digestion, the bolus is passed on to the stomach where the food matrix is further mixed and plant cells further macerated. Using in vitro studies, human mastication was determined to enhance the release of β-carotene from the plant matrix in one study by approximately 35 % during in vitro gastric and intestinal digestion. The particle size and the type of chewing had more impact on carotenoid bioaccessibility than cell wall presence, with smaller particle size and more fine chewing significantly enhancing bioaccessibility(Reference Low, D'Arcy and Gidley32), presumably due to enhanced access of digestion enzymes. In line with these results, emulsions with small droplet diameters (0·2 v. 23 µm) improved β-carotene transfer from lipid droplets to mixed micelles and bioaccessibility from approximately 35–60 %(Reference Salvia-Trujillo, Verkempinck and Sun33).

To a low extent, the acidic pH (3–5) of the stomach may lead to small losses of β-carotene, resulting in the formation of carotenoid-cations (CarH+) first(Reference Konovalov and Kispert34), which may then result in transcis isomerisation. However, most studies investigating digestion do not suggest that significant isomerisation takes place under physiological conditions(Reference Ferruzzi, Lumpkin and Schwartz35Reference Tyssandier, Reboul and Dumas37). Losses of β-carotene during digestion were reported from in vitro studies to range in the area of 30–70 %(Reference Blanquet-Diot, Soufi and Rambeau38, Reference Courraud, Berger and Cristol39), especially in the presence of oxidising compounds such as iron(Reference Kopec, Gleize and Borel40), likely resulting in the formation of β-apo-carotenals and epoxides(Reference Sy, Dangles and Borel41, Reference Sy, Dangles and Borel42). However, a recent clinical study using C13 β-carotene supports the conclusion of the only published clinical study dedicated to this topic, i.e. that β-carotene is fairly robust to human digestive conditions, with total losses <2 %(Reference Kopec, Caris-Veyrat and Nowicki43). Regarding host-produced enzymes, gastric lipase has been reported to be responsible for the digestion of approximately 5–40 % of ingested lipids(Reference Alminger, Aura and Bohn44), breaking down TAG, which are present together with carotenoids in lipid droplets. To which extent variations in gastric lipase concentrations translate into altered β-carotene bioaccessibility is not known. A practical limitation to study such effects is the non-availability of recombinant gastric-lipase or a suitable replacement which would allow studying effects more realistically in vitro (Reference Alminger, Aura and Bohn44).

In the small intestine, under the influence of bile acids, free fatty acids (FFAs), mono- and di-acylglycerols and phospholipids, originating mostly from the diet following digestion and under the influence of pancreatic lipase, lipid droplets are further processed to allow for mixed micelles formation of about 3–8 nm diameter(Reference Sy, Gleize and Dangles45, Reference Parker46). Proteins may aid in emulsification(Reference Soukoulis and Bohn47), but may also act as inhibitors thereof, preventing their transfer to mixed micelles(Reference Saini, Nile and Park48). A higher content of lipase and bile likely foster micellisation(Reference Bohn, Desmarchelier and Dragsted28, Reference Corte-Real, Desmarchelier and Borel49). Likewise, the presence of lipids in a meal rather foster micellisation(Reference Bohn50, Reference Borel51), while fibre(Reference Palafox-Carlos, Ayala-Zavala and Gonzalez-Aguilar52) or high minerals may hamper micelle formation(Reference Biehler, Hoffmann and Krause53, Reference Borel, Desmarchelier and Dumont54). The more apolar carotenoids, including β-carotene, are hypothesised to accumulate in core of the mixed micelle. The mixed micelles then diffuse through the mucus layer to the unstirred water layer of the enterocytes in the small intestine, where β-carotene may be taken up (Fig. 1).

Fig. 1. (Colour online) Processing of β-carotene during digestion. All factors that impinge on matrix-release, transfer from lipid droplets to mixed micelles, and their diffusion to the enterocyte surface can alter bioaccessibility and thus bioavailability of β-carotene. By contrast, the influence of the colon and its microbiota remains unclear.

Though micellisation of cis-isomers is higher than their all-trans forms(Reference Ferruzzi, Lumpkin and Schwartz35), cellular uptake appears higher for the all-trans-form(Reference During, Hussain and Morel55). Whether and to what extent cellular uptake depends on the mucus layer is still unclear, but the mucus layer does not appear to constitute a barrier to absorption(Reference Kaulmann, Andre and Schneider56). While β-carotene absorption was earlier thought to occur mostly by passive diffusion(Reference Hollander and Ruble57), the involvement of several apical membrane proteins (transporters), including scavenger-receptor class B-type I (SR-B1(Reference Sun58)), also known as SR-BI(Reference During, Dawson and Harrison59) as well as CD36(Reference During, Dawson and Harrison59Reference Borel, Lietz and Goncalves62) has meanwhile strongly been suggested, following studies with compounds inhibiting these proteins or in models over-expressing them. Genetic association studies have indeed found that SNP in SCARB1 (Reference Borel, Lietz and Goncalves62) and ABCA1 (Reference Borel, Desmarchelier and Nowicki25) were correlated with β-carotene bioavailability as determined by plasma and plasma-triglyceride-rich lipoproteins appearance, respectively (summarised in Supplementary table 1). However, it is not known whether these proteins have a direct or an indirect effect on β-carotene absorption. Indeed, recent results suggest that CD36 may indirectly modulate the apical to intracellular flux of β-carotene by modulating the synthesis rate of chylomicrons(Reference Buttet, Traynard and Tran63) in which β-carotene is incorporated. A recent candidate gene association study has suggested that ABCG5, and thus the heterodimer ABCG5/G8, could be involved in the efflux of a fraction of absorbed β-carotene back to the intestinal lumen(Reference Borel, Desmarchelier and Nowicki25). It is not clear if and to which extent Niemann-Pick disease, type C1, gene-like 1 or other membrane proteins participate in β-carotene absorption(Reference During, Dawson and Harrison59).

Finally, following gastro-intestinal digestion, a large proportion of non-absorbed β-carotene reaches the colon (as much as 50–95 %). It is unclear what happens under the influence of the gut microbiota(Reference Bohn64), but it has been shown that a large proportion of β-carotene is degraded into unknown compounds(Reference Kaulmann, Andre and Schneider56, Reference Goni, Serrano and Saura-Calixto65), as much as 98 % for pure β-carotene(Reference Serrano, Goni and Saura-Calixto66). However, a report by Mosele et al.(Reference Mosele, Macia and Romero67) points out to a high stability of β-carotene following in vitro colonic fermentation. The remainder, according to a report up to 83 %(Reference Shiau, Mobarhan and Stacewicz-Sapuntzakis68), is thus excreted in the faeces. It is likely that the matrix and microbiota differences add significantly to this variation. However, at present, there is no evidence that the colon plays a significant part in the absorption of β-carotene or its metabolites.

In summary, all factors that influence micellisation are likely to influence further β-carotene bioavailability and ADME aspects. The crucial part that micellisation plays for bioaccessibility is also well reflected by the high correlation between in vitro derived bioaccessibility and measures of bioavailability in vivo (Reference Tyssandier, Reboul and Dumas37, Reference Reboul, Richelle and Perrot69).

Intracellular metabolism and basolateral secretion of β-carotene by enterocytes

After having crossed the apical membrane, β-carotene must cross the polarised intestinal cell to be secreted at its basolateral side (Fig. 2A). Little is known about the intracellular transport and metabolism of β-carotene in the enterocyte. Nevertheless, since β-carotene is insoluble in water, intracellular binding protein(s) is/are likely to be involved(Reference Reboul and Borel70). This protein could be BCO1, which is mainly localised in the cytosol of mature enterocytes from the jejunum(Reference Duszka, Grolier and Azim71), as it is the main enzyme cleaving β-carotene(Reference Amengual, Widjaja-Adhi and Rodriguez-Santiago23, Reference Grolier, Duszka and Borel72Reference dela Sena, Narayanasamy and Riedl74), and because it has a great affinity for β-carotene. An intracellular transport protein could also be BCO2, although we assume that its mitochondrial localisation(Reference Amengual, Lobo and Golczak75) is not compatible with its involvement as an intracellular transporter of β-carotene. An intracellular β-carotene transporter could also be a fatty acid binding protein (FABP), more likely liver FABP (L-FABP / FABP1), which is also present in the intestine and displays high-affinity binding for various hydrophobic ligands(Reference Gajda and Storch76).

Fig. 2. (Colour online) (A) Candidate proteins for β-carotene metabolism within the enterocyte. When genetic variants have been associated with β-carotene bioavailability(Reference Borel, Desmarchelier and Nowicki25), the encoded proteins are coloured in grey. Dotted lines indicate regulations, i.e. regulation of BCO1 and SR-BI expression by ISX and regulation of chylomicron synthesis by SR-BI and CD36. (B) Candidate proteins that can modulate postprandial blood chylomicron β-carotene concentrations. When genetic variants have been associated with postprandial chylomicron β-carotene response to dietary β-carotene(Reference Borel, Desmarchelier and Nowicki25), the encoded proteins are coloured in grey. The dotted line indicates that this pathway is assumed but not demonstrated. (C) Proteins involved in the liver metabolism of β-carotene. Note that, to focus on β-carotene and for improved clarity, the fate of chylomicron retinyl esters in the liver is not shown, as well as the liver metabolism of retinol that involves numerous proteins(Reference Borel and Desmarchelier235). The liver is the hub of β-carotene metabolism: it is the main organ that stores β-carotene and distributes it to the peripheral tissues. β-Carotene reaches the liver mainly as β-carotene and retinyl esters, mainly RP, originating from β-carotene cleavage in the enterocyte and incorporated in chylomicrons. β-Carotene is then mostly stored in hepatic stellate cells. When genetic variants have been associated with blood β-carotene concentrations(Reference Hendrickson, Hazra and Chen80, Reference Yabuta, Urata and Wai Kun81, Reference Ferrucci, Perry and Matteini236), the encoded proteins are coloured in grey. βC: β-carotene, ABCA1: ATP binding cassette subfamily A member 1, ABCB1: ATP-binding cassette, sub-family B (MDR/TAP), member 1, ABCG5/G8: ATP-binding cassette, sub-family G member 5 and 8, ATRA: all-trans-retinoic acid, BCO1: β-carotene oxygenase 1, BCO2: β-carotene oxygenase 2, BCO2: β-carotene oxygenase 2, CD36: CD36 molecule, CXCL8: C-X-C motif chemokine ligand 8, ELOVL2: elongation of very long chain fatty acids protein 2, FABP: fatty acid binding protein, GPIHBP1: glycosylphosphatidylinositol-anchored high density lipoprotein binding protein 1, HL: hepatic lipase (encoded by LIPC), HSPGs: heparan sulphate proteoglycans, ISX: intestine specific homoeobox (transcription factor under the control of retinoic acid, regulating expression of SR-BI and BCO1), LDLR: LDL-receptor, LPL: lipoprotein lipase, LRP1: LDL-receptor-related protein 1, MTP: microsomal TAG transfer protein, NPC1L1: Niemann Pick C1-like 1, PKD1L2: polycystin 1-like 2 (gene/pseudogene), RBP4: serum retinol-binding protein, ROL: retinol, RP: retinyl palmitate and other retinyl esters coming from βC cleavage in the enterocyte, RPE65: retinal pigment epithelium-specific 65 kDa protein, SAR1B: secretion associated Ras-related GTPase 1B, SOD2: superoxide dismutase 2, SR-BI: scavenger receptor class B type I, TCF7L2: transcription factor 7-like 2, TTR: transthyretin.

At this step, it is important to emphasise that only a fraction of absorbed β-carotene is metabolised in the enterocyte. The importance of this fraction, which was estimated at about 70 % by using stable isotope methods(Reference Tang, Qin and Dolnikowski77), likely depends on the vitamin A status of the body (see the next section). The secretion mechanism of β-carotene at the basolateral side of the enterocyte likely depends on its centric cleavage by BCO1, producing retinal, which is then, following conversion to retinol, mainly re-esterified by lecithin-retinol acyl-transferase (LRAT). It is assumed that the parent molecule is incorporated in nascent chylomicrons(Reference Borel, Grolier and Mekki78), while the less apolar β-carotene metabolites, which are produced by eccentric cleavage by BCO2, are secreted in the portal blood. Indeed, one genome-wide association study(79) and two candidate gene association studies(Reference Hendrickson, Hazra and Chen80, Reference Yabuta, Urata and Wai Kun81) have shown that SNP in BCO1, the main β-carotene metabolising enzyme, were associated with blood plasma β-carotene concentration. Other gene-association studies involving SNP in BCO1 and postprandial β-carotene and retinyl palmitate responses(Reference Lietz, Oxley and Leung82, Reference Leung, Hessel and Meplan83) confirmed that this gene and its variants are key regulators of blood concentrations of these vitamin A forms.

The mechanisms responsible for the incorporation of β-carotene in chylomicrons are poorly understood. It is hypothesised that they involve enzymes/Apo responsible for the assembly of chylomicrons, e.g. microsomal TAG transfer protein (MTP), apoA-IV, secretion associated Ras-related GTPase 1B (SAR1B) and apoB48. A recent candidate gene association study(Reference Borel, Desmarchelier and Nowicki25) has also suggested that the protein involved in intestinal HDL secretion(Reference Brunham, Kruit and Iqbal84), i.e. ATP binding cassette subfamily A member 1 (ABCA1), may also be involved in β-carotene secretion in intestinal HDL.

Finally, we suggest that the secretion of β-carotene metabolites in the portal blood might involve basolateral membrane proteins that can aid in the efflux of these metabolites, e.g. ATP binding cassette subfamily B member 1 (ABCB1), which encodes for P-glycoprotein(Reference Harrison85). Furthermore, several candidate gene association studies have shown that genetic variants in elongation of very long chain fatty acids protein 2 (ELOVL2, also termed ELOVL fatty acid elongase 2) play a significant role in carotenoid absorption(Reference Borel, Desmarchelier and Nowicki24Reference Borel, Desmarchelier and Nowicki26). This is possibly due to the inhibitory effect of EPA, which is further elongated to docosapentaenoic acid and DHA by ELOVL2, on β-carotene absorption(Reference Mashurabad, Kondaiah and Palika86).

It is now acknowledged that vitamin A status can regulate β-carotene absorption and cleavage efficiency via a negative feedback loop: the higher the vitamin A status, the lower β-carotene absorption efficiency and cleavage, and inversely. The mechanism involves an intestinal transcription factor termed intestine specific homoeobox (ISX), which acts as a repressor of SCARB1 and BCO1 upon ATRA activation(Reference Lobo, Amengual and Baus87, Reference Lobo, Hessel and Eichinger88). Following vitamin A uptake, the intracellular concentrations of ATRA increase, inducing ISX expression. Consequently, less β-carotene is taken up by the enterocyte, and less β-carotene can be converted to retinal. When the intracellular concentration of ATRA drops, which is assumed to be the case when dietary vitamin A intake is low, ISX exerts less repressor activity towards SCARBI and BCO1 and consequently β-carotene uptake and conversion increase. A study in Zambian children with hypervitaminosis A supports this regulation. Indeed, these children had high serum carotenoid concentrations(Reference Mondloch, Gannon and Davis89) and many of them experienced hypercarotenodermia during mango season, a period of high provitamin A carotenoid intake. This might indicate as a possible explanation that conversion of provitamin A carotenoids to retinal by BCO1 was more inhibited by the hypervitaminosis A than their absorption via SR-BI, which is encoded by SCARB1. This is not surprising, as provitamin A carotenoid absorption involves not only SR-B1 but also CD36(Reference Borel, Lietz and Goncalves62), which is not assumed to be regulated by ISX. In a recent candidate genes association study a SNP in ISX together with SNP in other genes, was associated with the variability in β-carotene bioavailability(Reference Borel, Desmarchelier and Nowicki25). It was also reported in another study that a SNP in the ISX binding site in the BCO1 promoter (rs6564851) was associated with decreased conversion rates of β-carotene by 50 % and increased fasting blood concentrations of β-carotene(Reference Lobo, Amengual and Baus87). These associations support that genetic variations in this gene are key determinants of blood β-carotene concentrations.

In summary, it is assumed that the interplay between the β-carotene metabolising enzymes BCO1/2 and potential intracellular transporters possibly in conjunction with yet unidentified efflux transport proteins and those involved in chylomicron synthesis are among the most crucial actors influencing β-carotene bioavailability. Furthermore, the host vitamin A status which is detected by ISX and likely further mediated via retinoic acid signalling pathways also constitutes a paramount ‘critical control point’ in the metabolism of β-carotene.

Postprandial blood transport of newly absorbed β-carotene from the intestine to the liver

The enterocytes are assumed to secrete most of the newly absorbed β-carotene into chylomicrons, though it has been suggested that water-soluble β-carotene metabolites, e.g. apo-carotenals, could be secreted in the portal circulation and therefore directly reach the liver (Fig. 2B(Reference Harrison85)). In general, it is assumed that for compounds with a log P below approximately 5, portal absorption would predominate(Reference Charman and Stella90), which would be the case for some of the β-apo-carotenoids. Chylomicrons also contain retinyl esters, mainly retinyl palmitate(Reference Sauvant, Mekki and Charbonnier91), which originate either from esterification of retinol produced by the BCO1-mediated cleavage of β-carotene, or from re-esterification of preformed vitamin A present in the diet. It has been shown that most retinyl palmitate and β-carotene are not exchanged between lipoproteins and remain in chylomicrons and their remnants during their intravascular metabolism(Reference Blomhoff, Helgerud and Dueland92, Reference Tyssandier, Choubert and Grolier93). Thus, most β-carotene incorporated into chylomicron remnants, which are produced during vascular lipolysis of chylomicron TAG by both lipoprotein lipase (LPL) and glycosylphosphatidylinositol anchored high density lipoprotein binding protein 1 (GPIHBP1)(Reference Dallinga-Thie, Franssen and Mooij94), is taken up by hepatocytes during the postprandial period(Reference Blomhoff, Helgerud and Rasmussen95). This uptake involves several proteins, e.g. the LDL-receptor (LDLR), the LDL-receptor-related protein 1 (LRP1), SR-B1 and heparan sulphate proteoglycans (HSPGs)(Reference Dallinga-Thie, Franssen and Mooij94). Candidate gene association studies have also found that SNP in LPL (Reference Herbeth, Gueguen and Leroy96) were associated with blood β-carotene concentration.

The fact that β-carotene is carried in the blood by chylomicrons during the postprandial period implies that its metabolism is closely related to that of these TAG-rich lipoproteins. This is supported by a recent study that has shown that genetic variants in genes involved in chylomicron metabolism, i.e. transcription factor 7-like 2 (TCF7L2), ApoB, LIPC (which encodes hepatic lipase) and ABCA1 modulate the postprandial chylomicron β-carotene response to a meal that contained dietary β-carotene(Reference Borel, Desmarchelier and Nowicki25). Note also that a SNP in LIPC (Reference Borel, Moussa and Reboul97) was also associated with fasting blood β-carotene concentration.

In short, the transport from the intestine to the liver is mostly governed by proteins involved in chylomicron metabolism in the blood (e.g. ApoB and hepatic lipase) and likely also by proteins involved in chylomicron uptake by the liver (e.g. LRP1 and LDLR).

Liver metabolism and blood transport of β-carotene and its metabolites from the liver to extra-hepatic tissues

The liver is the main storage organ for vitamin A mainly in the form of retinyl esters. It has been estimated that for healthy, well-nourished individuals, approximately 70 % of vitamin A present in the body is stored in the liver(Reference O'Byrne and Blaner98). Following chylomicron-remnant uptake by the liver, which involves cell surface receptors (see the previous section), it is assumed that chylomicron remnant retinyl palmitate and β-carotene are released in hepatocytes during chylomicron remnant metabolism (Fig. 2C). They are then assumed to follow different metabolic pathways. Retinyl palmitate is assumed to be hydrolysed by a retinyl ester hydrolase to retinol. Retinol is then assumed to bind to cellular retinol-binding protein, type I (CRBPI / RBP1)(Reference Ong99) and to be transported to either the site where it is transferred to retinol-binding protein 4 (RBP4) or to hepatic stellate cells (also known as fat-storing cells, lipocytes or Ito cells) where it is re-esterified by LRAT(Reference Ong, MacDonald and Gubitosi100, Reference Rose101). Interestingly, hepatic LRAT expression is regulated by vitamin A status(Reference Blomhoff, Helgerud and Rasmussen95). This regulation likely involves ATRA and its respective response elements activated by the liganded nuclear hormone receptors RAR and RXR and further interaction with DNA. This regulation is proposed to give rise to a positive feedback loop when cellular ATRA concentrations are high, turning on hepatic stellate cell LRAT expression(Reference Nagatsuma, Hayashi and Hano102) and increasing the synthesis of retinyl esters(Reference O'Byrne and Blaner98) in these cells(Reference Wake103, Reference Wake104). These cells store approximately 70–90 % of liver vitamin A(Reference O'Byrne and Blaner98).

Contrarily to chylomicron retinyl palmitate, the fate of chylomicron β-carotene in the liver is barely known. How β-carotene is released from chylomicrons and how it is transported into hepatocytes remains unanswered. Concerning its cleavage, it is assumed that it is either cleaved to retinal by BCO1, which is highly expressed in hepatic stellate cells(Reference Shmarakov, Fleshman and D'Ambrosio105), or by BCO2, which is apparently more expressed in hepatocytes(Reference Shmarakov, Fleshman and D'Ambrosio105). The fraction of β-carotene that does not undergo this cleavage is either incorporated into VLDL and secreted into the blood, or stored in lipid droplets in parenchymal and hepatic stellate cells(Reference Shmarakov, Fleshman and D'Ambrosio105, Reference Lakshman, Asher and Attlesey106). The mechanism involved in the mobilisation of β-carotene stores is not known, but we hypothesise that it requires the hydrolysis of lipid droplet TAG.

The liver secretes vitamin A in the form of retinol either into the bile partly also as oxidised and/or further conjugated, i.e. glucuronidated metabolites(Reference Zachman and Olson107, Reference Zachman, Singer and Olson108), or directly into the blood. The liver secretes retinol into the blood arising partly from β-carotene metabolism but also as pro-vitamin A carotenoids, mainly β-carotene. Retinol is bound to serum retinol binding protein (RBP4), which in turn binds to transthyretin (TTR), stabilising the complex(Reference Peterson109). β-Carotene is incorporated in VLDL. Retinol associated with RBP4/TTR is taken up by two structurally related membrane receptors: stimulated by retinoic acid 6 (STRA6)(Reference Kawaguchi, Yu and Honda110) and the recently discovered STRA6-like receptor, also known as RBP4 receptor-2 (RBPR2)(Reference Alapatt, Guo and Komanetsky111). Retinol uptake via STRA6 depends on a functional coupling with intracellular LRAT(Reference Amengual, Golczak and Palczewski112). STRA6 and RBPR2 exhibit different tissue expression patterns: STRA6 is expressed in numerous tissues but not in the liver and intestine, where RBPR2 is mostly expressed(Reference Alapatt, Guo and Komanetsky111). VLDL-β-carotene and LDL-β-carotene, which originate from VLDL metabolism, are most likely taken up by tissues via LDL-receptor dependent mechanisms(Reference Thomas and Harrison113), requiring the tissue to express the LDL-receptor. However, candidate gene association studies have also found SNP associated with blood β-carotene concentration for SR-BI(Reference Borel, Moussa and Reboul114), also participating in the cellular uptake of HDL(Reference Hoekstra115), in addition to cellular uptake of β-carotene by enterocytes(Reference van Bennekum, Werder and Thuahnai60, Reference Borel, Lietz and Goncalves62, Reference Harrison85).

In summary, β-carotene can be stored mainly in the liver or alternative organs within the organism (and further cleaved by BCO1/2), secreted into the blood or into the bile. Bile excretion occurs via single or multiple oxidations at various locations of the derivative and further glucuronidation. Secretion into the bloodstream may follow targeted β-carotene cleavage at various locations or further incorporation and transport by VLDL.

β-Carotene metabolism, biodistribution, bioactivation and excretion

β-Carotene intake depends on the individual food intake in addition with an important influence of the individual food matrix and well as the food quality as outlined earlier. Further β-carotene metabolism includes isomerisation of the conjugated double bond system to various geometric isomers such as 9-cis- and 13-cis-β-carotene (Fig. 3). These cis-isomers are present in minor amounts in the raw food matrix(Reference Ben-Amotz and Fishier116Reference Vasquez-Caicedo, Sruamsiri and Fau-Carle118) and isomer concentration may increase also due to food processing including simple heat based cooking(Reference Lessin and Schwartz29). It may further be induced in the human organism by various processes(Reference Hieber, King and Fau-Morioka119Reference Relevy, Rühl and Harari121), where cis-isomers seem to be preferred v. all-trans-isomers. Whether this accumulation is due to favourable biophysical properties or by targeted and preferred uptake of cis-isomers is currently not known. For retinoids with a cis-isomeric structure, some enzymes were found which mediate targeted and non-targeted isomerisation, such as retinal pigment epithelium-specific 65 kDa protein(Reference Shyam, Gorusupudi and Nelson122, Reference Redmond, Poliakov and Kuo123) conversion to 11-cis-isomers, and Sphingolipid delta(4)-desaturase / desaturase 1 (DES1) (Reference Kaylor, Yuan and Cook124) to 9-cis- / 13-cis-isomers. In addition, binding proteins such as RBP1 and RLBP1 and retinol dehydrogenases RDH4/5 have specific cis-isomer preferred binding/transporter and synthesis mechanisms(Reference Parker and Crouch125Reference Huang, Possin and Saari127). If these targeted isomerisation and isomer-binding properties also exist for carotenoids and apo-carotenoids seems likely via the same or alternative enzymes and binding proteins but has not been studied so far.

Fig. 3. (Colour online) Metabolism of β-carotene with major metabolites formed in vivo. Involved enzymes, binding proteins, receptors and target genes involved in β-carotene metabolism towards bioactive retinoids. Derivatives marked with ‘*’ have been conclusively identified to be endogenously present. At – all-trans; SCARB1 – scavenger receptor class B type I; CD36 – cluster of density 36; ABCG5 / 8 – ATP binding cassette member 5 / 8; BCO1 – β-carotene oxygenase 1; BCO2 – β-carotene oxygenase 2; LRAT1 / 2 – lecithin retinol acyltransferase, DGAT1 / 2 – diacylglycerol O-acyltransferase 1 / 2; ISX – intestinal transcription factor; STRA6 – stimulated by retinoic acid 6; RBPR2 – retinol-binding protein receptor 2; RDH 5 / 10 – retinol dehydrogenase 5 / 10; DHRS3 / 9 – short-chain dehydrogenase/reductase 3 / 9; RBP1 / 2 / 4 – retinyl-binding protein 1 / 2 / 4; REH – retinyl-esterase; RETSAT – all-trans-retinol 13,14-reductase; ALDH1A1 / 2 / 3 – aldehyde dehydrogenase 1 family, member A1 / 2 / 3; CRABP1 / 2 – cellular-retinoic acid binding protein 1 / 2; RPE65 – retinal pigment epithelium-specific 65 kDa protein; DES1 – sphingolipid delta(4)-desaturase; RLBP1 – retinal-binding protein 1; RAR – retinoic acid receptor; RXR – retinoid-X receptor; TG2 – transglutaminase 2; SCD1 – stearoyl-CoA desaturase / (Δ−9-)desaturase-1; ELOVL6 – elongation of very long chain fatty acids protein 6; HOXB6 / 8 – homoeobox protein 6 / 8, HBEGF – heparin-binding-epidermal growth factor; RARRES2 – retinoic acid receptor responder protein 2 / chemerin; ADIPOQ – adiponectin; UCP1 – uncoupling protein 1, UGT2B7 – UDP-glucuronyltransferase-glucuronosyltransferase-2B7.

Following absorption, a large proportion of carotenoid and later retinoid metabolism is under control of nuclear hormone receptor signalling, as a partly auto-regulatory homoeostatic regulated process. Many steps, involving receptors such as RARβ, to anabolic enzymes including BCO1(Reference Bachmann, Desbarats and Pattison128), BCO2 and aldehyde dehydrogenase 1 family, member A3 (ALDH1A3), catabolic enzymes including CYP26A1 and LRAT and binding proteins, including RBP1, RBP4 and cellular-retinoic acid binding protein 2 (CRABP2) are under control of RAR-RXR- and PPAR-RXR-mediated signalling(Reference Gericke, Ittensohn and Mihaly129, Reference Balmer and Blomhoff130). RAR-RXR- and PPAR-RXR-mediated signalling also controls various other important lipid metabolic processes and places carotenoids as precursors of important regulators of general lipid metabolism as reviewed earlier(Reference Evans and Mangelsdorf131). This auto-regulation of general lipid metabolism and nuclear hormone-mediated signalling includes various target genes and several of these specific target genes are key for eliciting beneficial health effects of β-carotene, especially regarding cancer, allergic inflammatory disorders such as asthma and various CVD. As a consequence, carotenoid metabolism and usage towards bioactive retinoids for further bioactive signalling seems to be likely dependent on: (a) sufficient levels of available carotenoids in the human organism and (b) a targeted metabolic bioactivation pathway to elicit beneficial activities of carotenoids.

A long list of enzymes and binding proteins (summarised in Fig. 3) is responsible for the metabolism of carotenoids to bioactive retinoids in a temporal and spatial highly controlled manner. The initial steps are the cleavage of carotenoids by the two known human carotenoid-oxygenases BCO1 and 2(Reference Lindqvist, Sharvill and Sharvill132, Reference Wu, Guo and Wang133). The resulting apo-carotenals (named retinal in the case of apo-15′-carotenal), can then further be reduced to alcohols and esterified to store retinoids as retinyl esters, a reaction mediated by LRAT and diacylglycerol O-acyltransferase 1/2 (DGAT1/2)(Reference Orland, Anwar and Cromley134Reference Ross and Zolfaghari137). Retinyl-esters can further be de-esterified by esterases (REHs) to alcohols and especially retinol to serve as precursors for later bioactivation(Reference Schreiber, Taschler and Preiss-Landl138, Reference Schreiber, Taschler and Wolinski139). Retinol can later be transported by various retinol binding proteins and further be oxidised in target tissues by retinol dehydrogenases, mainly RDH4/5 and 10 as well as short-chain dehydrogenase / reductase 3 and 9 (DHRS 3 and 9) (Reference Napoli140, Reference Kumar, Sandell and Trainor141). Inter- and intra-cellular transport of various forms of retinals and retinols is mediated by specific binding proteins including RBP1, 2 and 4 and RLBP1(Reference Napoli142). The physiological and nutritional relevance of additional apo-carotenals and apo-carotenoic acids remains unclear, and they were described as low affinity activators / competitive antagonists of nuclear hormone receptors, as reviewed previously(Reference Eroglu and Harrison143). Unfortunately, a clear link between the low physiological and nutritional relevant levels in human subjects or high level and further dependent biological-mediated signalling were not described yet, and thus no nutritional or physiological relevance can currently clearly be associated with these derivatives.

The major intermediate bioactive precursor is retinal, the visual pigment, which can be obtained possibly from multiple sources: (a) direct cleavage of carotenoids via BCO1(Reference Lindqvist and Andersson144) or (b) via a physiological unclear and indirect cascade via BCO2(Reference Amengual, Widjaja-Adhi and Rodriguez-Santiago23) or (c) via oxidation from retinol(Reference Napoli145). Retinol is present at low levels in the food matrix and in highly homoeostatically regulated levels in blood. However, retinal becomes also significantly available via cleavage of retinyl-esters, the major dietary relevant and major storage form, which can be cleaved to retinol and oxidised to retinal in target tissues(Reference Napoli142, Reference D'Ambrosio, Clugston and Blaner146). The oxidation of retinal to ATRA is the key step to yield the lipid-hormone ATRA. Retinoic acid dehydrogenases (aldehyde dehydrogenase 1 family, member A1, A2 and A3), in association with cellular retinoic acid binding proteins 1 and 2 are strictly controlling this bio-activation(Reference D'Ambrosio, Clugston and Blaner146). All-trans retinoic acid (ATRA) can further initiate, via ligand activation of the RARα, β, γ, transcriptional signalling of a large array of target genes(Reference Balmer and Blomhoff130) (Fig. 5) including: TG2 – transglutaminase 2; HBEGF – heparin-binding-epidermal growth factor; RARRES2 – RAR responder protein 2 / chemerin; ADIPOQ – adiponectin; UCP1 – uncoupling protein 1, as well as the afore-mentioned auto-regulated targets in the retinoid metabolism cascade.

While ATRA as the endogenous RAR-ligand is well accepted(Reference Petkovich147), the existence of an endogenous RXR-ligand has largely been mysterious(Reference de Lera, Krezel and Rühl148, Reference Calleja, Messaddeq and Chapellier149). The ATRA isomer 9-cis-retinoic acid (9CRA) has been described to be ‘the’ endogenous ligand for this crucial heterodimer nuclear hormone receptor(Reference Allenby, Bocquel and Saunders150Reference Heyman, Mangelsdorf and Dyck152). Unfortunately, the endogenous, physiological and nutritional relevance of 9CRA has been questioned and remained unclear(Reference Rühl, Krzyzosiak and Niewiadomska-Cimicka153Reference Schmidt, Brouwer and Nau159). Recently, the lipid hormone 9-cis-13,14-dihydroretinoic acid (9CDHRA) has been identified as the endogenous and physiological relevant RXR ligand(Reference de Lera, Krezel and Rühl148, Reference Rühl, Krzyzosiak and Niewiadomska-Cimicka153). Further examinations about the nutritional relevance are currently under investigation(Reference Rühl, Krezel and de Lera160). 9CDHRA can activate RXR and via this route also initiate non-permissive heterodimers, such as RXR-PPAR, RXR-liver X receptors (LXRs), RXR-farnesoid receptors (FXR) and RXR-nuclear hormone receptor 4A (NR4A) protein, involved in the expression of a wide array of genes involved in inflammation and lipid metabolism, as reviewed in Desvergne(Reference Desvergne161). A large overlap was found between beneficial anti-inflammatory effects of carotenoids in general and lipid metabolism for many RXR-heterodimer-mediated signalling pathway targets(Reference Evans and Mangelsdorf131, Reference Desvergne161Reference Mangelsdorf, Ong and Dyck166).

The precise metabolic pathway leading to 9CDHRA is not known yet and likely involves retinol-saturase (RETSAT(Reference Rühl, Krzyzosiak and Niewiadomska-Cimicka153, Reference Moise, Kuksa and Imanishi167), and / or DES-1) and the binding protein RBP1(Reference Kaylor, Yuan and Cook124, Reference Rühl, Krzyzosiak and Niewiadomska-Cimicka153), as well as retinaldehyde binding protein 1 (RLBP1 / CRALBP(Reference Saari, Huang and Possin168, Reference Huang, Jarjour and Oumesmar169)). Additionally, novel still non-identified carotenoids are speculated to be more direct precursors for 9CDHRA (postulated and also outlined in Fig. 5).

Other retinoids such as 13-cis- or 9,13-dicis-RA(Reference Allenby, Bocquel and Saunders150, Reference Horst, Reinhardt and Goff170, Reference Chen, Sass and Seltmann171), all-trans-13,14-dihydroretinoic acid(Reference Vahlquist172, Reference Torma, Asselineau and Andersson173) and phase 1 metabolites including 4-hydroxy- or 4-oxo-retinoic acid, or other hydroxyl- / oxo-metabolites are likely of minor importance for RAR- or RXR-mediated signalling processes(Reference Schmidt, Brouwer and Nau159, Reference Pijnappel, Hendriks and Folkers174Reference Baron, Heise and Blaner177). Alternative phase 1 metabolism may also occur for β-carotene and would result in epoxidation and hydroxylation to produce a large array of epoxy-, hydroxyl- and oxo-carotenoids (examples shown in Fig. 3, reviewed earlier(Reference Bohm and Bitsch178)). These hydroxy- / oxo-retinoids can further be conjugated via phase 2 metabolism by UDP-glucuronyltransferase-glucosyltransferase to yield water-soluble excretion metabolites such as retinyl-, retinoic acid (retinoyl) and oxo-retinoic acid glucuronides, which were found in serum, faeces and urine(Reference Sass, Masgrau and Saurat179Reference Samokyszyn, Gall and Zawada181).

BCO1 as a critical bottleneck for the cleavage of β-carotene and enabling RAR- / RXR-mediated signalling

Human subjects centrally cleave β-carotene to retinal and following oxidation and reduction a larger array of multi-functional retinoids are created which have been detected endogenously (Fig. 3(Reference Böhm, Borel and Corte-Real13)). This BCO1-mediated conversion of β-carotene to retinal is therefore an important bottleneck, which is highly controlled and mediated by various factors: (A) availability of the substrate and saturation of the enzymatic conversion potential, (B) presence and relative levels of food derived inhibitors, (C) spatial and temporal regulation and localisation of the enzyme, (D) sex-specific regulations, (E) feedback regulations by bioactive products transcriptionally controlling BCO1-expression and (F) the previously reported polymorphisms of BCO1 as well as assisting proteins.

Starting with the overall conversion in the human organism, β-carotene was reported to vary largely in absorption efficiency (30–70 %) upon intestinal uptake, as explained earlier. This variation is in part due to the variation in BCO1 cleavage potency, partly explainable by frequently occurring polymorphisms of BCO1(Reference Lietz, Oxley and Leung82, Reference Leung, Hessel and Meplan83). Alternatively, many factors which are summarised here describe the inter- and auto-regulatory pathways in BCO1-mediated cleavage to retinoids, enabling RAR- or / and RXR-mediated signalling.

The main question is how BCO1 and its mediated cleavage to centric cleavage metabolites are regulated in the human organism. As described earlier, six major steps (A–F) have been reported and identified. First, the availability of the substrate is usually a major factor for increased conversion and resulting product levels (Fig. 4A). This conversion was presented as a saturating curve, plateauing at β-carotene levels in the range of 15 000–40 000 and 80 000–240 000 nm for β-cryptoxanthin(Reference Lindqvist and Andersson144). For comparison, the endogenous levels for β-carotene were in average range of 360 nm in serum and up to 31 830 nm in organs, while being highest in the adrenals and β-cryptoxanthin in the average range of 230 nm in serum and with highest tissue levels of 2900 nm in adrenals, as reviewed recently(Reference Böhm, Borel and Corte-Real13). We can thus conclude that these active ranges were not reached in serum, while tissue levels approach the saturation of enzyme conversion(Reference Böhm, Borel and Corte-Real13, Reference Lindqvist and Andersson144). It should be noted that serum and tissue levels do not represent freely available carotenoids, but mainly carotenoids attached to binding proteins and carotenoids associated in lipid vesicles in the membranes and lipid accumulating vesicles such as in the adipose tissue(Reference Böhm, Borel and Corte-Real13). In one study examining children(Reference Rühl, Taner and Schweigert182Reference Rühl184), a direct correlation between all-trans-β-carotene (ATβC) and ATRA serum levels was observed, resulting in a linear correlation with r 0·68, while in adults(Reference Mihaly, Marosvolgyi and Szegedi185Reference Lucas, Mihály and Lowe187) no such correlation was observed (r −0·14), likely due to auto-regulative processes. The highest relevant level of ATβC in serum of children was 756 nm (Fig. 4B) and 1628 ng/ml (3031 nm) in adults (Fig. 4C), displaying no plateau for conversion to ATRA, based on endogenous relevant β-carotene levels (Fig. 4B and C).

Fig. 4. (Colour online) BCO1 localisation and metabolic properties. (A) In vitro kinetic analysis of purified recombinant human BCO1 with β-carotene and β-cryptoxanthin, as published earlier from Lindqvist and Andersson(Reference Lindqvist and Andersson144). (B) Direct correlation newly calculated based on of serum ATβC to ATRA in children in Germany with different ethical backgrounds(Reference Gruber, Taner and Mihaly183, Reference Rühl184). (C) Direct correlation based on serum ATβC to ATRA levels in Hungarian adults (n 40, Lucas et al.(Reference Lucas, Mihály and Lowe187)) This figure is just present in the original study in ng/ml, while 1 ng/ml ATRA corresponds to 3.3 nM and 1 ng/ml ATβC to 1.86 nM. (D) Distribution of BCO1 mRNA expression in human tissues, as published previously in Lindqvist and Andersson(Reference Lindqvist and Andersson144) (PBL – peripheral blood lymphocytes). (E) Differentially expressed genes and pathways by β-carotene v. control diet. Gene expression analysis of different tissues on a control diet supplemented with βC v. control diet (containing adequate vitamin A)-fed mice. A description of the mouse study can be found in van Helden et al.(Reference van Helden, Godschalk and Swarts14). The global transcriptome data were extracted from Gene Expression Omnibus (GEO, Superseries GSE98847), containing lung (GSE98845), liver (GSE98846) and inguinal white adipose tissue (iWAT; GSE27271) and were normalised per tissue and genotype with both sexes included for comparison between sexes. Sex-specific number of differential expressed genes (P < 0·05) are given in number and fold change (FC) of males v. females. (F) ATRA levels in serum (nm) and lung ((pmol/ml / 10−2 m) of control treated (CTRL), low-β-carotene (βC)-diet supplemented (LBC) and high-βC supplemented ferrets (HBC) adapted from Liu et al.(Reference Liu, Wang and Bronson237). Panels A, B, D and F are adapted from van Helden et al.(Reference van Helden, Godschalk and Swarts14) and Lindqvist and Andersson(Reference Lindqvist and Andersson144) and were permitted to reproduction under copyright.

The second modification factor are other carotenoids such as canthaxanthin, lutein and zeaxanthin(Reference Grolier, Duszka and Borel72, Reference van Vliet, van Schaik and Schreurs188), which can inhibit BCO1-mediated conversion partly in a competitive manner. The specific mechanisms of these phenomena were not investigated deeper, and neither have nutritional relevant ranges and relevant ratios been examined. It is likely that these three carotenoids can attach and bind to the active site of the BCO1, thereby inhibiting the binding and enzymatic conversion of known pro-vitamin A carotenoid BCO1 substrates (β-carotene, α-carotene, β-cryptoxanthin, apo-8′-carotenal and lycopene). In in vitro studies, it was reported that three times higher levels of lutein (compared with β-carotene) interfered with β-carotene conversion. These ratios reflect relevant physiological conditions(Reference van Vliet, van Schaik and Schreurs188), which can be obtained after targeted lutein/zeaxanthin supplementation or dietary intake of fruits and vegetables high in these carotenoids, as reviewed recently(Reference Böhm, Borel and Corte-Real13). Additionally assisting proteins such as RBP1 and RBP2, acting as intracellular sensors of endogenous retinoid status are thereby important contributors for ATβC conversion to ATRA(Reference Lietz, Lange and Rimbach189Reference Nagao191).

The third modifying factor is the specific spatial and temporal regulation of BCO1 expression in the human organism(Reference Lindqvist and Andersson144). Highest BCO1 expression was found in different parts of the intestine, with highest expression levels in the jejunum (Fig. 4D). Other relevant tissues are reproduction organs testis and prostate in males as well as ovaries in females, comparable with levels found in kidney, liver, skeletal muscle and stomach (Fig. 4D). Not displayed in this figure are the relatively high expression levels observed in the eye(192).

As a fourth modification, sex specific regulations were observed. In male mice and rats, a connection between testosterone and carotenoid as well as BCO1-expression was found(Reference Ford, Moran and Smith193Reference Campbell, Stroud and Nakamura195), while oestrogen / testosterone correlated in older woman with carotenoid levels(Reference Maggio, de Vita and Lauretani196). If non-reproduction-related organs also display this regulation, depending on sexual steroid hormones, was not further investigated. One indicator are higher ATRA and lower retinol serum levels in women v. men(Reference Soderlund, Sjoberg and Svard197), likely as a consequence of higher levels of β-carotene in their serum / plasma, mainly due to the less healthy nutritional status(Reference Böhm, Borel and Corte-Real13, Reference El-Sohemy, Baylin and Kabagambe198, Reference Tucker, Chen and Vogel199), or higher BCO1 presence and activity.

Recently, also gene expression microarray studies were conducted, using wild type (WT) and BCO1-knockout male and female mice, on a low but sufficient vitamin A diet with or without additional β-carotene supplementation(Reference van Helden, Godschalk and Swarts14, Reference van Helden, Godschalk and von Lintig200Reference van Helden, Heil and van Schooten202), which were used to provide further insights into the differential effects of dietary β-carotene supplementation. It was observed that β-carotene supplementation alters only a small number of overlapping genes in the lung of wild-type male and female mice (n 20, Fig. 4E), while a larger number of the genes altered by β-carotene supplementation were sex specific, n 631 in female and n 306 in male mice(Reference van Helden, Godschalk and Swarts14) (Fig. 4E). This difference was even more striking in the BCO1 knockout mice, where 1433 genes were specifically affected in females and 1385 in males, with only 89 being affected in both sexes and, strikingly, for 85 of these, the direction of expression was oppositely regulated between the sexes (Fig. 4E). The number of genes affected by β-carotene in the liver of BCO1 knockout mice was less than that of the lung (Fig. 4E).

In inguinal white adipose tissue (iWAT) a different pattern emerged with a large number of genes being specifically regulated upon β-carotene exposure in WT female mice (4840) (Fig. 4E). The number of genes specifically regulated by β-carotene exposure in males of WT (276) and BCO1 knockout (1168) genotype was comparable between iWAT v. lung, while BCO1 knockout females showed a reduced number (567 v. 1433) in iWAT. This was strikingly also the case for the much larger number (33 WT v. 130 KO) of common genes in WT mice, and is likely explained by the important role of WAT in steroid hormone metabolism, especially for oestrogens and progestogens in females(Reference DiSilvestro, Petrosino and Aldoori203Reference Bonet, Canas and Ribot205).

Detailed analysis of the effects of dietary β-carotene supplementation identified a strong down-regulation of RXRα, as well as the pro-adipogenesis trigger PPARγ and its target genes in iWAT of WT female mice(Reference Amengual, Gouranton and van Helden206). This effect was likely dependent on BCO1-mediated retinoid production, since it was not observed in BCO1 knockout female mice and this effect was associated with a reduction in WAT mass, resulting in a reduced adiposity index. This adiposity lowering effect is in line with observations that show that oral ATRA administration induces energy expenditure and fat mass lowering in mice, with WAT being one of the contributing tissues(Reference Mercader, Ribot and Murano207).

The fifth modifying effect is the regulation of BCO1 on the transcriptional level. A feedback mechanism was identified, partly already before a clear identification, characterisation and expression of BCO1 in mice, rats and chicken and human subjects(Reference Lietz, Lange and Rimbach189, Reference Fierce, de Morais Vieira and Piantedosi190, Reference Parvin and Sivakumar208Reference van Vliet, van Vlissingen and van Schaik210). The conversion and ratio of ATβC to retinoids, especially all-trans-retinal, was used to identify BCO1 activity(Reference Bachmann, Desbarats and Pattison128, Reference Wang, Tang and Fox211). Feedback mechanisms were claimed as a direct involvement of ATRA-RAR-interaction and transcriptional modification of BCO1 expression was shown. ATRA–RAR-mediated signalling is suggested to regulate BCO1 expression identified, either indirectly by retinal conversion per homogenate ratio or directly by mRNA quantification, as a negative feedback mechanism(Reference Bachmann, Desbarats and Pattison128). Treatments of rats with ATRA, retinyl acetate, β-carotene or a synthetic RAR-agonist (Ro41-5253) significantly reduced BCO1 activity identified by retinal conversion(Reference Bachmann, Desbarats and Pattison128). Focusing on ATRA, retinyl acetate and β-carotene treatments to rats, it was found that also serum retinoic acid levels increased and partly negatively correlated with reduced intestinal BCO1 activity(Reference Bachmann, Desbarats and Pattison128). This mechanism was thereby identified as an important negative feedback regulation for retinoid and mainly RAR-mediated signalling. It is noteworthy that nutritional supplementation with high β-carotene can result even in decreased local levels of ATRA with potential negative effects and increased vulnerability towards carcinogenesis as shown in β-carotene-supplemented ferrets (Fig. 4F(Reference Wang, Liu and Bronson15)). This highlights the limits of β-carotene-signalling mediated autoregulation using non-nutritional relevant to high β-carotene stimuli, with even previously reported negative side effects in human subjects as found in the ATBC and Carotene and Retinol Efficacy Trial (CARET) studies(Reference Omenn, Goodman and Thornquist11, 12). That high retinoid stimuli can induce negative side effects was recently described in mice(Reference Rubin, Ross and Stephensen212Reference Garcia, Rühl and Herz215), and seems also relevant for high-nutritional relevant β-carotene supplementation in ferrets and human subjects(Reference Rühl, Taner and Schweigert182Reference Rühl184, Reference Melhus, Michaelsson and Kindmark216Reference Boelsma, van de Vijver and Goldbohm219). In consequence, low / moderate β-carotene supplementation seems to be a tolerable nutritional stimulus to which the mammalian / human organism can respond. Recently it was reported that glucocorticoid regulated pathways and hepatocyte nuclear factor (HNF)1α and HNF4α pathways are important regulators of BCO1 expression(Reference Yamaguchi, Sunto and Goda220).

In addition, RXR-PPARα and -PPARγ-mediated signalling was identified as an alternative mechanism, providing positive feedback(Reference Gong, Tsai and Yan221, Reference Gong, Marisiddaiah and Rubin222). The PPARα and PPARγ nuclear hormone receptor heterodimers can be either activated by an RXR-ligand or alternatively by the respective PPAR ligand. For PPARs and HNF4α FFAs and fatty acid metabolites have been identified as natural ligands (Fig. 5(Reference Schupp and Lazar223, Reference Dhe-Paganon, Duda and Iwamoto224)). After a high-dietary intake of fat this important regulatory pathway is initiated by increased levels of FFAs as a direct result of the diet rich in fat and results further in increased BCO1-expression as a direct feedback to this high-fat diet. These two nuclear hormone receptor heterodimers need either an RXR-ligand as well as / or a PPAR-ligand. Dietary transglutaminases can provide PPAR ligands, thus synchronising fat, and concomitantly carotenoid, uptake / availability with BCO1 up-regulation. It was described that the main BCO1-metabolite ATRA is regulating via ATRA–RAR-mediated signalling various important steps in lipid metabolism(Reference Amengual, Gouranton and van Helden206, Reference Kim, Zuccaro and Costabile225, Reference Landrier, Kasiri and Karkeni226), with a focus also on counteracting fat accumulation via energy dissipation in adipose tissue(Reference Landrier, Kasiri and Karkeni226Reference Bonet, Ribot and Palou229) and regulation of insulin secretion(Reference Takeda, Sriram and Chan230, Reference Brun, Grijalva and Rausch231). The second possibility for negative feedback is the potential synthesis of the endogenous RXR-ligand 9CDHRA, starting from still non-identified carotenoid precursors (Rühl et al., unpublished results(Reference Rühl, Krezel and de Lera160)). This means that the endogenous RAR ligand ATRA and the endogenous RXR-ligand 9CDHRA obtain potential opposite regulation on their own synthesis via positive or negative feedback control mechanism of BCO1 expression and further activity (Fig. 5). As a consequence, levels and dietary intake of specific carotenoid precursors may influence or even control BCO1-mediated synthesis of endogenous RAR- or RXR-ligands and further controlling metabolic processes associated with lipid metabolism with relevance for obesity and diabetes.

Fig. 5. Transcriptional regulation of BCO1 metabolism and affected biological processes. Schematic summary of metabolism of the endogenous RAR-activator ATRA starting from ATβC, via all-trans-retinal (ATRAL) to ATRA, which can further activate RAR-RXR-mediated transcriptional signalling. In parallel the newly identified endogenous RXR-ligand 9-cis-13,14-dihydroretinoic acid (9CDHRA) can be created starting from putative carotenoid via putative retinal-analogues to 9CDHRA, which can further activate RXR-hepatocyte nuclear factor (HNF)4α, -PPAR α or -PPARγ-mediated transcriptional signalling. These three receptors (HNF4α, PPARα and PPARγ) can be activated by their ligands, free fatty acids (FFAs) and other metabolites originating from fatty acids. The RAR- or RXR-mediated signalling can positively or negatively alter transcriptional regulated BCO1-expression. LUT, lutein; CAN, canthaxanthin; ZEA, zeaxanthin.

In summary, BCO1 represents a bottleneck for β-carotene-conversion to bioactive retinoids and further RAR-mediated transcriptional activation and signalling. A food matrix high in different natural occurring carotenoids leads to metabolism from pro-vitamin A carotenoids (β-carotene and β-cryptoxanthin) to retinoids, while additional carotenoids such as lutein, zeaxanthin and canthaxanthin may inhibit this conversion. This balanced carotenoid mixture, which is mainly present in fruits and vegetables as a balanced diet, is likely resulting in a much lower conversion to retinoic acids, than an equimolar supplementation with β-carotene as a nutritional supplement. In addition, a diet high in fat induces a strong activation of PPARα-, PPARγ-RXR / HNF4α-mediated signalling, following increased BCO1-expression and increased ATRA levels. Furthermore, this ATRA can induce increased RAR-RXR-mediated signalling as a natural feedback mechanism to stimulate lipid catabolism and blocking fat accumulation. In conclusion, a balanced diet rich in carotenoids originating from fruits and vegetables or alternatively a balanced carotenoid supplementation as present in fruit and vegetable extracts or to be developed ‘smart’ carotenoid supplements and not artificial single high carotenoid supplementation should result in moderate retinoid synthesis with the potential of balancing fat accumulation and stimulating fat usage in the human organism.

Conclusions

β-Carotene availability from the diet, as well as BCO1-mediated cleavage towards bioactive ATRA under consideration of tissue and sex dependent regulation are the two main bottlenecks for enabling retinoid-mediated signalling, as the most well-known processes of β-carotene's metabolic action. Serum β-carotene levels are affected by various aspects concerning our diet and human polymorphisms. Further bioactive signalling starting from serum and tissue ATRA levels is highly homoeostatically auto-regulated by various mechanisms, including nutritional stimuli and sex hormonal regulatory pathways. A strong deficiency of β-carotene in the diet or strong nutritional / supplemental β-carotene stimuli can result in altered retinoic acid levels but without any reported significantly altered further biological-mediated signalling (reviewed in Böhm et al.(Reference Böhm, Borel and Corte-Real13) and Watzl et al.(Reference Watzl, Bub and Brandstetter232Reference Watzl, Bub and Briviba234)).

How short-term or long-term β-carotene or alternative stimuli affecting BCO1-cleavage and further can alter RAR- / RXR-mediated signalling must be further evaluated in human supplementation trials additionally examining known or postulated β-carotene-dependent health-biomarkers especially including novel omics-based disease marker.

As a final conclusion, the human organism seems to have a high flexibility balancing high and low dietary β-carotene availability based on a complex homoeostatic regulation for maintaining physiological crucial RAR-mediated signalling. The optimal dietary range of β-carotene in concert with other nutrients is highlighted in an additional review(Reference Böhm, Borel and Corte-Real13).

Supplementary material

The supplementary material for this article can be found at https://doi.org/10.1017/S0029665118002641.

Acknowledgements

The insights from EU-COST action POSITIVE (FA 1403) and EUROCAROTEN (CA 15136) are much appreciated.

Financial support

None.

Conflict of interest

The authors declare no conflict of interest.

Authorship

T. B., C. D., S. N. E., J. K., E. v. S., R. R. and P. B. contributed in setting up and writing the manuscript. RR was the invited speaker at the Glasgow symposium.

References

1.Biehler, E, Alkerwi, A, Hoffmann, L et al. (2012) Contribution of violaxanthin, neoxanthin, phytoene and phytofluene to total carotenoid intake: assessment in Luxembourg. J Food Comp Anal 25, 5665.Google Scholar
2.Pezdirc, K, Hutchesson, MJ, Williams, RL et al. (2016) Consuming high-carotenoid fruit and vegetables influences skin yellowness and plasma carotenoids in young women: a single-blind randomized crossover trial. J Acad Nutr Diet 116, 12571265.Google Scholar
3.Wawrzyniak, A, Hamulka, J, Friberg, E et al. (2013) Dietary, anthropometric, and lifestyle correlates of serum carotenoids in postmenopausal women. Eur J Nutr 52, 19191926.Google Scholar
4.Fraser, GE, Jaceldo-Siegl, K, Henning, SM et al. (2016) Biomarkers of dietary intake are correlated with corresponding measures from repeated dietary recalls and food-frequency questionnaires in the adventist health study-2. J Nutr 146, 586594.Google Scholar
5.Sluijs, I, Cadier, E, Beulens, JW et al. (2015) Dietary intake of carotenoids and risk of type 2 diabetes. Nutr Metab Cardiovasc Dis 25, 376381.Google Scholar
6.Sugiura, M, Nakamura, M, Ogawa, K et al. (2015) High-serum carotenoids associated with lower risk for developing type 2 diabetes among Japanese subjects: Mikkabi cohort study. BMJ Open Diabetes Res Care 3, e000147.Google Scholar
7.Wang, Y, Chung, SJ, McCullough, ML et al. (2014) Dietary carotenoids are associated with cardiovascular disease risk biomarkers mediated by serum carotenoid concentrations. J Nutr 144, 10671074.Google Scholar
8.Karppi, J, Kurl, S, Ronkainen, K et al. (2013) Serum carotenoids reduce progression of early atherosclerosis in the carotid artery wall among Eastern Finnish men. PLoS ONE 8, e64107.Google Scholar
9.Bjelakovic, G, Nikolova, D, Gluud, LL et al. (2007) Mortality in randomized trials of antioxidant supplements for primary and secondary prevention: systematic review and meta-analysis. JAMA 297, 842857.Google Scholar
10.Bjelakovic, G, Nikolova, D, Gluud, LL et al. (2008) Antioxidant supplements for prevention of mortality in healthy participants and patients with various diseases. Cochrane Database Syst Rev 14, CD007176.Google Scholar
11.Omenn, GS, Goodman, GE, Thornquist, MD et al. (1996) Effects of a combination of beta carotene and vitamin A on lung cancer and cardiovascular disease. New Engl J Med 334, 11501155.Google Scholar
12.The Alpha-tocopherol Beta-carotene Cancer Prevention Group (1994) The effect of vitamin E and beta carotene on the incidence of lung cancer and other cancers in male smokers. The Alpha-Tocopherol, Beta Carotene Cancer Prevention Study Group. N Engl J Med 330, 10291035.Google Scholar
13.Böhm, V, Borel, P, Corte-Real, J et al. (2018) From carotenoid intake to carotenoid/metabolite blood and tissue concentrations – implications for dietary intake recommendations. Nutr Rev submitted.Google Scholar
14.van Helden, YG, Godschalk, RW, Swarts, HJ et al. (2011) Beta-carotene affects gene expression in lungs of male and female Bcmo1 (-/-) mice in opposite directions. Cell Mol Life Sci 68, 489504.Google Scholar
15.Wang, XD, Liu, C, Bronson, RT et al. (1999) Retinoid signaling and activator protein-1 expression in ferrets given beta-carotene supplements and exposed to tobacco smoke. J Natl Cancer Inst 91, 6066.Google Scholar
16.Krinsky, NI & Johnson, EJ (2005) Carotenoid actions and their relation to health and disease. Mol Aspects Med 26, 459516.Google Scholar
17.Krinsky, NI & Yeum, KJ (2003) Carotenoid-radical interactions. Biochem Biophys Res Commun 305, 754760.Google Scholar
18.Piga, R, van Dartel, D, Bunschoten, A et al. (2014) Role of frizzled6 in the molecular mechanism of beta-carotene action in the lung. Toxicology 320, 6773.Google Scholar
19.Kaulmann, A & Bohn, T (2014) Carotenoids, inflammation, and oxidative stress – implications of cellular signaling pathways and relation to chronic disease prevention. Nutr Res 34, 907929.Google Scholar
20.Rühl, R, Bub, A & Watzl, B (2008) Modulation of plasma all-trans retinoic acid concentrations by the consumption of carotenoid-rich vegetables. Nutrition 24, 12241226.Google Scholar
21.Prakash, P, Liu, C, Hu, KQ et al. (2004) Beta-carotene and beta-apo-14′-carotenoic acid prevent the reduction of retinoic acid receptor beta in benzo[a]pyrene-treated normal human bronchial epithelial cells. J Nutr 134, 667673.Google Scholar
22.Ben-Dor, A, Nahum, A, Danilenko, M et al. (2001) Effects of acyclo-retinoic acid and lycopene on activation of the retinoic acid receptor and proliferation of mammary cancer cells. Arch Biochem Biophys 391, 295302.Google Scholar
23.Amengual, J, Widjaja-Adhi, MA, Rodriguez-Santiago, S et al. (2013) Two carotenoid oxygenases contribute to mammalian provitamin A metabolism. J Biol Chem 288, 3408134096.Google Scholar
24.Borel, P, Desmarchelier, C, Nowicki, M et al. (2015) Lycopene bioavailability is associated with a combination of genetic variants. Free Radic Biol Med 83, 238244.Google Scholar
25.Borel, P, Desmarchelier, C, Nowicki, M et al. (2015) A combination of single-nucleotide polymorphisms is associated with interindividual variability in dietary beta-carotene bioavailability in healthy men. J Nutr 145, 17401747.Google Scholar
26.Borel, P, Desmarchelier, C, Nowicki, M et al. (2014) Interindividual variability of lutein bioavailability in healthy men: characterization, genetic variants involved, and relation with fasting plasma lutein concentration. Am J Clin Nutr 100, 168175.Google Scholar
27.Desmarchelier, C & Borel, P (2017) Overview of carotenoid bioavailability determinants: from dietary factors to host genetic variations. Trends Food Sci Technol in press, doi: org/10.1016/j.tifs.2017.1003.1002.Google Scholar
28.Bohn, T, Desmarchelier, C, Dragsted, LO et al. (2017) Host-related factors explaining interindividual variability of carotenoid bioavailability and tissue concentrations in humans. Mol Nutr Food Res 61, 1600685.Google Scholar
29.Lessin, WJ & Schwartz, SJ (1997) Quantification of cistrans isomers of provitamin A carotenoids in fresh and processed fruits and vegetables. J Agric Food Chem 45, 37283732.Google Scholar
30.Marx, M, Stuparic, M, Schieber, A et al. (2003) Effects of thermal processing on transcis-isomerization of β-carotene in carrot juices and carotene-containing preparations. Food Chem 83, 609617.Google Scholar
31.Cilla, A, Bosch, L, Barberá, R et al. (2018) Effect of processing on the bioaccessibility of bioactive compounds – a review focusing on carotenoids, minerals, ascorbic acid, tocopherols and polyphenols. J Food Compost Anal 68, 315.Google Scholar
32.Low, DY, D'Arcy, B & Gidley, MJ (2015) Mastication effects on carotenoid bioaccessibility from mango fruit tissue. Food Res Int 67, 238246.Google Scholar
33.Salvia-Trujillo, L, Verkempinck, SH, Sun, L et al. (2017) Lipid digestion, micelle formation and carotenoid bioaccessibility kinetics: influence of emulsion droplet size. Food Chem 229, 653662.Google Scholar
34.Konovalov, V & Kispert, LD (1999) AM1, INDO/S and optical studies of carbocations of carotenoid molecules. Acid induced isomerization. J Chem Soc 2, 901910.Google Scholar
35.Ferruzzi, MG, Lumpkin, JL, Schwartz, SJ et al. (2006) Digestive stability, micellarization, and uptake of beta-carotene isomers by Caco-2 human intestinal cells. J Agric Food Chem 54, 27802785.Google Scholar
36.Failla, ML, Chitchumroonchokchai, C & Ishida, BK (2008) In vitro micellarization and intestinal cell uptake of cis isomers of lycopene exceed those of all-trans lycopene. J Nutr 138, 482486.Google Scholar
37.Tyssandier, V, Reboul, E, Dumas, JF et al. (2003) Processing of vegetable-borne carotenoids in the human stomach and duodenum. Am J Physiol Gastrointest Liver Physiol 284, G913G923.Google Scholar
38.Blanquet-Diot, S, Soufi, M, Rambeau, M et al. (2009) Digestive stability of xanthophylls exceeds that of carotenes as studied in a dynamic in vitro gastrointestinal system. J Nutr 139, 876883.Google Scholar
39.Courraud, J, Berger, J, Cristol, JP et al. (2013) Stability and bioaccessibility of different forms of carotenoids and vitamin A during in vitro digestion. Food Chem 136, 871877.Google Scholar
40.Kopec, RE, Gleize, B, Borel, P et al. (2017) Are lutein, lycopene, and beta-carotene lost through the digestive process? Food Funct 8, 14941503.Google Scholar
41.Sy, C, Dangles, O, Borel, P et al. (2013) Iron-induced oxidation of (all-E)-beta-carotene under model gastric conditions: kinetics, products, and mechanism. Free Radical Biol Med 63, 195206.Google Scholar
42.Sy, C, Dangles, O, Borel, P et al. (2015) Stability of bacterial carotenoids in the presence of iron in a model of the gastric compartment – comparison with dietary reference carotenoids. Arch Biochem Biophys 572, 89100.Google Scholar
43.Kopec, RE, Caris-Veyrat, C, Nowicki, M et al. (2018) Production of asymmetric oxidative metabolites of 13C β-carotene during digestion in the gastrointestinal lumen of healthy men. Am J Clin Nutr 108, 803813.Google Scholar
44.Alminger, M, Aura, AM, Bohn, T et al. (2014) In vitro models for studying secondary plant metabolite digestion and bioaccessibility. Comp Rev Food Sci Food Saf 13, 413436.Google Scholar
45.Sy, C, Gleize, B, Dangles, O et al. (2012) Effects of physicochemical properties of carotenoids on their bioaccessibility, intestinal cell uptake, and blood and tissue concentrations. Mol Nutr Food Res 56, 13851397.Google Scholar
46.Parker, RS (1996) Absorption, metabolism, and transport of carotenoids. FASEB J 10, 542551.Google Scholar
47.Soukoulis, C & Bohn, T (2018) A comprehensive overview on the micro- and nano-technological encapsulation advances for enhancing the chemical stability and bioavailability of carotenoids. Crit Rev Food Sci Nutr 58, 136.Google Scholar
48.Saini, RK, Nile, SH & Park, SW (2015) Carotenoids from fruits and vegetables: chemistry, analysis, occurrence, bioavailability and biological activities. Food Res Int 76, Part 3, 735750.Google Scholar
49.Corte-Real, J, Desmarchelier, C, Borel, P et al. (2017) Magnesium affects spinach carotenoid bioaccessibility in vitro depending on intestinal bile and pancreatic enzyme concentrations. Food Chem 239, 751759.Google Scholar
50.Bohn, T (2008) Bioavailability of non-provitamin A carotenoids. Curr Nutr Food Sci 4, 240258.Google Scholar
51.Borel, P (2003) Factors affecting intestinal absorption of highly lipophilic food microconstituents (fat-soluble vitamins, carotenoids and phytosterols). Clin Chem Lab Med 41, 979994.Google Scholar
52.Palafox-Carlos, H, Ayala-Zavala, JF & Gonzalez-Aguilar, GA (2011) The role of dietary fiber in the bioaccessibility and bioavailability of fruit and vegetable antioxidants. J Food Sci 76, R6R15.Google Scholar
53.Biehler, E, Hoffmann, L, Krause, E et al. (2011) Divalent minerals decrease micellarization and uptake of carotenoids and digestion products into Caco-2 cells. J Nutr 141, 17691776.Google Scholar
54.Borel, P, Desmarchelier, C, Dumont, U et al. (2016) Dietary calcium impairs tomato lycopene bioavailability in healthy humans. Br J Nutr 116, 20912096.Google Scholar
55.During, A, Hussain, MM, Morel, DW et al. (2002) Carotenoid uptake and secretion by CaCo-2 cells: beta-carotene isomer selectivity and carotenoid interactions. J Lipid Res 43, 10861095.Google Scholar
56.Kaulmann, A, Andre, CM, Schneider, YJ et al. (2015) Carotenoid and polyphenol bioaccessibility and cellular uptake from plum and cabbage varieties. Food Chem 197, 325332.Google Scholar
57.Hollander, D & Ruble, PE Jr (1978) Beta-carotene intestinal absorption: bile, fatty acid, pH, and flow rate effects on transport. Am J Physiol 235, E686E691.Google Scholar
58.Sun, H (2012) Membrane receptors and transporters involved in the function and transport of vitamin A and its derivatives. Biochim Biophys Acta 1821, 99112.Google Scholar
59.During, A, Dawson, HD & Harrison, EH (2005) Carotenoid transport is decreased and expression of the lipid transporters SR-BI, NPC1L1, and ABCA1 is downregulated in Caco-2 cells treated with ezetimibe. J Nutr 135, 23052312.Google Scholar
60.van Bennekum, A, Werder, M, Thuahnai, ST et al. (2005) Class B scavenger receptor-mediated intestinal absorption of dietary beta-carotene and cholesterol. Biochemistry (Mosc) 44, 45174525.Google Scholar
61.During, A, Doraiswamy, S & Harrison, EH (2008) Xanthophylls are preferentially taken up compared with beta-carotene by retinal cells via a SRBI-dependent mechanism. J Lipid Res 49, 17151724.Google Scholar
62.Borel, P, Lietz, G, Goncalves, A et al. (2013) CD36 and SR-BI are involved in cellular uptake of provitamin A carotenoids by Caco-2 and HEK cells, and some of their genetic variants are associated with plasma concentrations of these micronutrients in humans. J Nutr 143, 448456.Google Scholar
63.Buttet, M, Traynard, V, Tran, TT et al. (2014) From fatty-acid sensing to chylomicron synthesis: role of intestinal lipid-binding proteins. Biochimie 96, 3747.Google Scholar
64.Bohn, T (2016) Bioactivity of carotenoids – chasms of knowledge. Int J Vitam Nutr Res 10, 15.Google Scholar
65.Goni, I, Serrano, J & Saura-Calixto, F (2006) Bioaccessibility of beta-carotene, lutein, and lycopene from fruits and vegetables. J Agric Food Chem 54, 53825387.Google Scholar
66.Serrano, J, Goni, I & Saura-Calixto, F (2005) Determination of beta-carotene and lutein available from green leafy vegetables by an in vitro digestion and colonic fermentation method. J Agric Food Chem 53, 29362940.Google Scholar
67.Mosele, JI, Macia, A, Romero, MP et al. (2016) Stability and metabolism of Arbutus unedo bioactive compounds (phenolics and antioxidants) under in vitro digestion and colonic fermentation. Food Chem 201, 120130.Google Scholar
68.Shiau, A, Mobarhan, S, Stacewicz-Sapuntzakis, M et al. (1994) Assessment of the intestinal retention of beta-carotene in humans. J Am Coll Nutr 13, 369375.Google Scholar
69.Reboul, E, Richelle, M, Perrot, E et al. (2006) Bioaccessibility of carotenoids and vitamin E from their main dietary sources. J Agric Food Chem 54, 87498755.Google Scholar
70.Reboul, E & Borel, P (2011) Proteins involved in uptake, intracellular transport and basolateral secretion of fat-soluble vitamins and carotenoids by mammalian enterocytes. Prog Lipid Res 50, 388402.Google Scholar
71.Duszka, C, Grolier, P, Azim, EM et al. (1996) Rat intestinal beta-carotene dioxygenase activity is located primarily in the cytosol of mature jejunal enterocytes. J Nutr 126, 25502556.Google Scholar
72.Grolier, P, Duszka, C, Borel, P et al. (1997) In vitro and in vivo inhibition of beta-carotene dioxygenase activity by canthaxanthin in rat intestine. Arch Biochem Biophys 348, 233238.Google Scholar
73.dela Sena, C, Riedl, KM, Narayanasamy, S et al. (2014) The human enzyme that converts dietary provitamin A carotenoids to vitamin A is a dioxygenase. J Biol Chem 289, 1366113666.Google Scholar
74.dela Sena, C, Narayanasamy, S, Riedl, KM et al. (2013) Substrate specificity of purified recombinant human beta-carotene 15,15′-oxygenase (BCO1). J Biol Chem 288, 3709437103.Google Scholar
75.Amengual, J, Lobo, GP, Golczak, M et al. (2011) A mitochondrial enzyme degrades carotenoids and protects against oxidative stress. FASEB J 25, 948959.Google Scholar
76.Gajda, AM & Storch, J (2015) Enterocyte fatty acid-binding proteins (FABPs): different functions of liver and intestinal FABPs in the intestine. Prostaglandins Leukot Essent Fatty Acids 93, 916.Google Scholar
77.Tang, G, Qin, J, Dolnikowski, GG et al. (2003) Short-term (intestinal) and long-term (postintestinal) conversion of beta-carotene to retinol in adults as assessed by a stable-isotope reference method. Am J Clin Nutr 78, 259266.Google Scholar
78.Borel, P, Grolier, P, Mekki, N et al. (1998) Low and high responders to pharmacological doses of beta-carotene: proportion in the population, mechanisms involved and consequences on beta-carotene metabolism. J Lipid Res 39, 22502260.Google Scholar
79.WHO (1976) Vitamin A deficiency and xerophthalmia. Report of a Joint WHO/USAID Meeting. World Health Organ Tech Rep Ser, 188.Google Scholar
80.Hendrickson, SJ, Hazra, A, Chen, C et al. (2012) Beta-Carotene 15,15′-monooxygenase 1 single nucleotide polymorphisms in relation to plasma carotenoid and retinol concentrations in women of European descent. Am J Clin Nutr 96, 13791389.Google Scholar
81.Yabuta, S, Urata, M, Wai Kun, RY et al. (2016) Common SNP rs6564851 in the BCO1 gene affects the circulating levels of beta-carotene and the daily intake of carotenoids in healthy Japanese women. PLoS ONE 11, e0168857.Google Scholar
82.Lietz, G, Oxley, A, Leung, W et al. (2012) Single nucleotide polymorphisms upstream from the beta-carotene 15,15’-monooxygenase gene influence provitamin A conversion efficiency in female volunteers. J Nutr 142, 161s165s.Google Scholar
83.Leung, WC, Hessel, S, Meplan, C et al. (2009) Two common single nucleotide polymorphisms in the gene encoding beta-carotene 15,15′-monooxygenase alter beta-carotene metabolism in female volunteers. FASEB J 23, 10411053.Google Scholar
84.Brunham, LR, Kruit, JK, Iqbal, J et al. (2006) Intestinal ABCA1 directly contributes to HDL biogenesis in vivo. J Clin Invest 116, 10521062.Google Scholar
85.Harrison, EH (2012) Mechanisms involved in the intestinal absorption of dietary vitamin A and provitamin A carotenoids. Biochim Biophys Acta 1821, 7077.Google Scholar
86.Mashurabad, PC, Kondaiah, P, Palika, R et al. (2016) Eicosapentaenoic acid inhibits intestinal beta-carotene absorption by downregulation of lipid transporter expression via PPAR-alpha dependent mechanism. Arch Biochem Biophys 590, 118124.Google Scholar
87.Lobo, GP, Amengual, J, Baus, D et al. (2013) Genetics and diet regulate vitamin A production via the homeobox transcription factor ISX. J Biol Chem 288, 90179027.Google Scholar
88.Lobo, GP, Hessel, S, Eichinger, A et al. (2010) ISX is a retinoic acid-sensitive gatekeeper that controls intestinal beta,beta-carotene absorption and vitamin A production. FASEB J 24, 16561666.Google Scholar
89.Mondloch, S, Gannon, BM, Davis, CR et al. (2015) High provitamin A carotenoid serum concentrations, elevated retinyl esters, and saturated retinol-binding protein in Zambian preschool children are consistent with the presence of high liver vitamin A stores. Am J Clin Nutr 102, 497504.Google Scholar
90.Charman, WNA & Stella, VJ (1986) Effects of lipid class and lipid vehicle volume on the intestinal lymphatic transport of DDT. Int J Pharm 33, 165172.Google Scholar
91.Sauvant, P, Mekki, N, Charbonnier, M et al. (2003) Amounts and types of fatty acids in meals affect the pattern of retinoids secreted in human chylomicrons after a high-dose preformed vitamin A intake. Metabolism 52, 514519.Google Scholar
92.Blomhoff, R, Helgerud, P, Dueland, S et al. (1984) Lymphatic absorption and transport of retinol and vitamin D-3 from rat intestine. Evidence for different pathways. Biochim Biophys Acta 772, 109116.Google Scholar
93.Tyssandier, V, Choubert, G, Grolier, P et al. (2002) Carotenoids, mostly the xanthophylls, exchange between plasma lipoproteins. Int J Vitam Nutr Res 72, 300308.Google Scholar
94.Dallinga-Thie, GM, Franssen, R, Mooij, HL et al. (2010) The metabolism of triglyceride-rich lipoproteins revisited: new players, new insight. Atherosclerosis 211, 18.Google Scholar
95.Blomhoff, R, Helgerud, P, Rasmussen, M et al. (1982) In vivo uptake of chylomicron [3H]retinyl ester by rat liver: evidence for retinol transfer from parenchymal to nonparenchymal cells. Proc Natl Acad Sci USA 79, 73267330.Google Scholar
96.Herbeth, B, Gueguen, S, Leroy, P et al. (2007) The lipoprotein lipase serine 447 stop polymorphism is associated with altered serum carotenoid concentrations in the Stanislas Family Study. J Am Coll Nutr 26, 655662.Google Scholar
97.Borel, P, Moussa, M, Reboul, E et al. (2009) Human fasting plasma concentrations of vitamin E and carotenoids, and their association with genetic variants in apo C-III, cholesteryl ester transfer protein, hepatic lipase, intestinal fatty acid binding protein and microsomal triacylglycerol transfer protein. Br J Nutr 101, 680687.Google Scholar
98.O'Byrne, SM & Blaner, WS (2013) Retinol and retinyl esters: biochemistry and physiology. J Lipid Res 54, 17311743.Google Scholar
99.Ong, DE (1982) Purification and partial characterization of cellular retinol-binding protein from human liver. Cancer Res 42, 10331037.Google Scholar
100.Ong, DE, MacDonald, PN & Gubitosi, AM (1988) Esterification of retinol in rat liver. Possible participation by cellular retinol-binding protein and cellular retinol-binding protein II. J Biol Chem 263, 57895796.Google Scholar
101.Rose, AC (1982) Retinol esterification by rat liver microsomes. J Biol Chem 257, 24532459.Google Scholar
102.Nagatsuma, K, Hayashi, Y, Hano, H et al. (2009) Lecithin: retinol acyltransferase protein is distributed in both hepatic stellate cells and endothelial cells of normal rodent and human liver. Liver Int 29, 4754.Google Scholar
103.Wake, K (1974) Development of vitamin A-rich lipid droplets in multivesicular bodies of rat liver stellate cells. J Cell Biol 63, 683691.Google Scholar
104.Wake, K (1980) Perisinusoidal stellate cells (fat-storing cells, interstitial cells, lipocytes), their related structure in and around the liver sinusoids, and vitamin A-storing cells in extrahepatic organs. Int Rev Cytol 66, 303353.Google Scholar
105.Shmarakov, I, Fleshman, MK, D'Ambrosio, DN et al. (2010) Hepatic stellate cells are an important cellular site for beta-carotene conversion to retinoid. Arch Biochem Biophys 504, 310.Google Scholar
106.Lakshman, MR, Asher, KA, Attlesey, MG et al. (1989) Absorption, storage, and distribution of beta-carotene in normal and beta-carotene-fed rats: roles of parenchymal and stellate cells. J Lipid Res 30, 15451550.Google Scholar
107.Zachman, RD & Olson, JA (1964) Formation and enterohepatic circulation of water-soluble metabolites of retinol (vitamin A) in the rat. Nature 201, 12221223.Google Scholar
108.Zachman, RD, Singer, MB & Olson, JA (1966) Biliary secretion of metabolites of retinol and of retinoic acid in the guinea pig and chick. J Nutr 88, 137142.Google Scholar
109.Peterson, PA (1971) Characteristics of a vitamin A-transporting protein complex occurring in human serum. J Biol Chem 246, 3443.Google Scholar
110.Kawaguchi, R, Yu, J, Honda, J et al. (2007) A membrane receptor for retinol binding protein mediates cellular uptake of vitamin A. Science 315, 820825.Google Scholar
111.Alapatt, P, Guo, F, Komanetsky, SM et al. (2013) Liver retinol transporter and receptor for serum retinol-binding protein (RBP4). J Biol Chem 288, 12501265.Google Scholar
112.Amengual, J, Golczak, M, Palczewski, K et al. (2012) Lecithin: retinol acyltransferase is critical for cellular uptake of vitamin A from serum retinol-binding protein. J Biol Chem 287, 2421624227.Google Scholar
113.Thomas, SE & Harrison, EH (2016) Mechanisms of selective delivery of xanthophylls to retinal pigment epithelial cells by human lipoproteins. J Lipid Res 57, 18651878.Google Scholar
114.Borel, P, Moussa, M, Reboul, E et al. (2007) Human plasma levels of vitamin E and carotenoids are associated with genetic polymorphisms in genes involved in lipid metabolism. J Nutr 137, 26532659.Google Scholar
115.Hoekstra, M (2017) SR-BI as target in atherosclerosis and cardiovascular disease – a comprehensive appraisal of the cellular functions of SR-BI in physiology and disease. Atherosclerosis 258, 153161.Google Scholar
116.Ben-Amotz, A & Fishier, R (1998) Analysis of carotenoids with emphasis on 9-cis β-carotene in vegetables and fruits commonly consumed in Israel. Food Chem 62, 515520.Google Scholar
117.Khoo, HE, Prasad, KN, Kong, KW et al. (2011) Carotenoids and their isomers: color pigments in fruits and vegetables. Molecules 16, 17101738.Google Scholar
118.Vasquez-Caicedo, AL, Sruamsiri, P, Fau-Carle, R et al. (2005) Accumulation of all-trans-beta-carotene and its 9-cis and 13-cis stereoisomers during postharvest ripening of nine Thai mango cultivars. J Agric Food Chem 53, 48274835.Google Scholar
119.Hieber, AD, King, TJ, Fau-Morioka, S et al. (2000) Comparative effects of all-trans beta-carotene v. 9-cis beta-carotene on carcinogen-induced neoplastic transformation and connexin 43 expression in murine 10T1/2 cells and on the differentiation of human keratinocytes. Nutr Cancer 37, 234244.Google Scholar
120.You, CS, Parker, RS, Goodman, KJ et al. (1996) Evidence of cis-trans isomerization of 9-cis-beta-carotene during absorption in humans. Am J Clin Nutr 64, 177183.Google Scholar
121.Relevy, N, Rühl, R, Harari, A et al. (2015) 9-cis-Beta-carotene inhibits atherosclerosis development in female LDLR-/-mice. Funct Foods Health Dis 5, 6779.Google Scholar
122.Shyam, R, Gorusupudi, A, Nelson, K et al. (2017) RPE65 has an additional function as the lutein to meso-zeaxanthin isomerase in the vertebrate eye. Proc Natl Acad Sci USA 114, 1088210887.Google Scholar
123.Redmond, TM, Poliakov, E, Kuo, S et al. (2010) RPE65, visual cycle retinol isomerase, is not inherently 11-cis-specific: support for a carbocation mechanism of retinol isomerization. J Biol Chem 285, 19191927.Google Scholar
124.Kaylor, JJ, Yuan, Q, Cook, J et al. (2013) Identification of DES1 as a vitamin A isomerase in Muller glial cells of the retina. Nat Chem Biol 9, 3036.Google Scholar
125.Parker, RO & Crouch, RK (2010) Retinol dehydrogenases (RDHs) in the visual cycle. Exp Eye Res 91, 788792.Google Scholar
126.Kane, MA (2012) Analysis, occurrence, and function of 9-cis-retinoic acid. Biochim Biophys Acta 1821, 1020.Google Scholar
127.Huang, J, Possin, DE & Saari, JC (2009) Localizations of visual cycle components in retinal pigment epithelium. Mol Vis 15, 223234.Google Scholar
128.Bachmann, H, Desbarats, A, Pattison, P et al. (2002) Feedback regulation of beta,beta-carotene 15,15′-monooxygenase by retinoic acid in rats and chickens. J Nutr 132, 36163622.Google Scholar
129.Gericke, J, Ittensohn, J, Mihaly, J et al. (2013) Regulation of retinoid-mediated signaling involved in skin homeostasis by RAR and RXR agonists/antagonists in mouse skin. PLoS ONE 8, e62643.Google Scholar
130.Balmer, JE & Blomhoff, R (2002) Gene expression regulation by retinoic acid. J Lipid Res 43, 17731808.Google Scholar
131.Evans, RM & Mangelsdorf, DJ (2014) Nuclear receptors, RXR, and the big bang. Cell 157, 255266.Google Scholar
132.Lindqvist, A, Sharvill, J, Sharvill, DE et al. (2007) Loss-of-function mutation in carotenoid 15,15-monooxygenase identified in a patient with hypercarotenemia and hypovitaminosis A. J Nutr 137, 23462350.Google Scholar
133.Wu, L, Guo, X, Wang, W et al. (2016) Molecular aspects of beta, beta-carotene-9′,10′-oxygenase 2 in carotenoid metabolism and diseases. Exp Biol Med (Maywood) 241, 18791887.Google Scholar
134.Orland, MD, Anwar, K, Cromley, D et al. (2005) Acyl coenzyme A dependent retinol esterification by acyl coenzyme A: diacylglycerol acyltransferase 1. Biochim Biophys Acta 1737, 7682.Google Scholar
135.Jiang, W & Napoli, JL (2012) Reorganization of cellular retinol-binding protein type 1 and lecithin: retinol acyltransferase during retinyl ester biosynthesis. Biochim Biophys Acta 1820, 859869.Google Scholar
136.Chelstowska, S, Widjaja-Adhi, MA, Silvaroli, JA et al. (2016) Molecular basis for vitamin A uptake and storage in vertebrates. Nutrients 8(11), pii: E676.Google Scholar
137.Ross, AC & Zolfaghari, R (2004) Regulation of hepatic retinol metabolism: perspectives from studies on vitamin A status. J Nutr 134, 269s275s.Google Scholar
138.Schreiber, R, Taschler, U, Preiss-Landl, K et al. (2012) Retinyl ester hydrolases and their roles in vitamin A homeostasis. Biochim Biophys Acta 1821, 113123.Google Scholar
139.Schreiber, R, Taschler, U, Wolinski, H et al. (2009) Esterase 22 and beta-glucuronidase hydrolyze retinoids in mouse liver. J Lipid Res 50, 25142523.Google Scholar
140.Napoli, JL (2012) Physiological insights into all-trans-retinoic acid biosynthesis. Biochim Biophys Acta 1821, 152167.Google Scholar
141.Kumar, S, Sandell, LL, Trainor, PA et al. (2012) Alcohol and aldehyde dehydrogenases: retinoid metabolic effects in mouse knockout models. Biochim Biophys Acta 1821, 198205.Google Scholar
142.Napoli, JL (2016) Functions of intracellular retinoid binding-proteins. Subcell Biochem 81, 2176.Google Scholar
143.Eroglu, A & Harrison, EH (2013) Carotenoid metabolism in mammals, including man: formation, occurrence, and function of apocarotenoids. J Lipid Res 54, 17191730.Google Scholar
144.Lindqvist, A & Andersson, S (2002) Biochemical properties of purified recombinant human beta-carotene 15,15′-monooxygenase. J Biol Chem 277, 2394223948.Google Scholar
145.Napoli, JL (2017) Cellular retinoid binding-proteins, CRBP, CRABP, FABP5: effects on retinoid metabolism, function and related diseases. Pharmacol Ther 173, 1933.Google Scholar
146.D'Ambrosio, DN, Clugston, RD & Blaner, WS (2011) Vitamin A metabolism: an update. Nutrients 3, 63103.Google Scholar
147.Petkovich, M (1992) Regulation of gene expression by vitamin A: the role of nuclear retinoic acid receptors. Annu Rev Nutr 12, 443471.Google Scholar
148.de Lera, AR, Krezel, W & Rühl, R (2016) An endogenous mammalian retinoid X receptor ligand, at last! ChemMedChem 11, 10271037.Google Scholar
149.Calleja, C, Messaddeq, N, Chapellier, B et al. (2006) Genetic and pharmacological evidence that a retinoic acid cannot be the RXR-activating ligand in mouse epidermis keratinocytes. Genes Dev 20, 15251538.Google Scholar
150.Allenby, G, Bocquel, MT, Saunders, M et al. (1993) Retinoic acid receptors and retinoid X receptors: interactions with endogenous retinoic acids. Proc Natl Acad Sci USA 90, 3034.Google Scholar
151.Levin, AA, Sturzenbecker, LJ, Kazmer, S et al. (1992) 9-cis retinoic acid stereoisomer binds and activates the nuclear receptor RXR alpha. Nature 355, 359361.Google Scholar
152.Heyman, RA, Mangelsdorf, DJ, Dyck, JA et al. (1992) 9-cis retinoic acid is a high affinity ligand for the retinoid X receptor. Cell 68, 397406.Google Scholar
153.Rühl, R, Krzyzosiak, A, Niewiadomska-Cimicka, A et al. (2015) 9-cis-13,14-dihydroretinoic acid is an endogenous retinoid acting as RXR ligand in mice. PLoS Genet 11, e1005213.Google Scholar
154.Kiss, I, Rühl, R, Szegezdi, E et al. (2008) Retinoid receptor-activating ligands are produced within the mouse thymus during postnatal development. Eur J Immunol 38, 147155.Google Scholar
155.Rühl, R (2006) Method to determine 4-oxo-retinoic acids, retinoic acids and retinol in serum and cell extracts by liquid chromatography/diode-array detection atmospheric pressure chemical ionisation tandem mass spectrometry. Rapid Commun Mass Spectrom 20, 24972504.Google Scholar
156.Blomhoff, R & Blomhoff, HK (2006) Overview of retinoid metabolism and function. J Neurobiol 66, 606630.Google Scholar
157.Gundersen, TE, Bastani, NE & Blomhoff, R (2007) Quantitative high-throughput determination of endogenous retinoids in human plasma using triple-stage liquid chromatography/tandem mass spectrometry. Rapid Commun Mass Spectrom 21, 11761186.Google Scholar
158.Gundersen, TE (2006) Methods for detecting and identifying retinoids in tissue. J Neurobiol 66, 631644.Google Scholar
159.Schmidt, CK, Brouwer, A & Nau, H (2003) Chromatographic analysis of endogenous retinoids in tissues and serum. Anal Biochem 315, 3648.Google Scholar
160.Rühl, R, Krezel, W & de Lera, AR (2018) 9-Cis-13,14-dihydroretinoic acid, a new endogenous mammalian ligand of retinoid X receptor and the active ligand of a potential new vitamin A category: vitamin A5. Nutr Rev 76, 929941.Google Scholar
161.Desvergne, B (2007) RXR: from partnership to leadership in metabolic regulations. Vitam Horm 75, 132.Google Scholar
162.Rühl, R & Landrier, JF (2016) Dietary regulation of adiponectin by direct and indirect lipid activators of nuclear hormone receptors. Mol Nutr Food Res 60, 175184.Google Scholar
163.Mihaly, J, Gericke, J, Lucas, R et al. (2016) TSLP expression in the skin is mediated via RARgamma-RXR pathways. Immunobiology 221, 161165.Google Scholar
164.Roszer, T, Menendez-Gutierrez, MP, Cedenilla, M et al. (2013) Retinoid X receptors in macrophage biology. Trends Endocrinol Metab 24, 460468.Google Scholar
165.Szanto, A, Narkar, V, Shen, Q et al. (2004) Retinoid X receptors: X-ploring their (patho)physiological functions. Cell Death Differ 11 (Suppl 2), S126S143.Google Scholar
166.Mangelsdorf, DJ, Ong, ES, Dyck, JA et al. (1990) Nuclear receptor that identifies a novel retinoic acid response pathway. Nature 345, 224229.Google Scholar
167.Moise, AR, Kuksa, V, Imanishi, Y et al. (2004) Identification of all-trans-retinol:all-trans-13,14-dihydroretinol saturase. J Biol Chem 279, 5023050242.Google Scholar
168.Saari, JC, Huang, J, Possin, DE et al. (1997) Cellular retinaldehyde-binding protein is expressed by oligodendrocytes in optic nerve and brain. Glia 21, 259268.Google Scholar
169.Huang, JK, Jarjour, AA, Oumesmar, BN et al. (2011) Retinoid X receptor gamma signaling accelerates CNS remyelination. Nat Neurosci 14, 4553.Google Scholar
170.Horst, RL, Reinhardt, TA, Goff, JP et al. (1995) Identification of 9-cis,13-cis-retinoic acid as a major circulating retinoid in plasma. Biochemistry 34, 12031209.Google Scholar
171.Chen, WC, Sass, JO, Seltmann, H et al. (2000) Biological effects and metabolism of 9-cis-retinoic acid and its metabolite 9,13-di-cis-retinoic acid in HaCaT keratinocytes in vitro: comparison with all-trans-retinoic acid. Arch Dermatol Res 292, 612620.Google Scholar
172.Vahlquist, A (1982) Vitamin A in human skin: I. detection and identification of retinoids in normal epidermis. J Invest Dermatol 79, 8993.Google Scholar
173.Torma, H, Asselineau, D, Andersson, E et al. (1994) Biologic activities of retinoic acid and 3,4-didehydroretinoic acid in human keratinocytes are similar and correlate with receptor affinities and transactivation properties. J Invest Dermatol 102, 4954.Google Scholar
174.Pijnappel, WW, Hendriks, HF, Folkers, GE et al. (1993) The retinoid ligand 4-oxo-retinoic acid is a highly active modulator of positional specification. Nature 366, 340344.Google Scholar
175.Niederreither, K, Abu-Abed, S, Schuhbaur, B et al. (2002) Genetic evidence that oxidative derivatives of retinoic acid are not involved in retinoid signaling during mouse development. Nat Genet 31, 8488.Google Scholar
176.Schmidt, CK, Volland, J, Hamscher, G et al. (2002) Characterization of a new endogenous vitamin A metabolite. Biochim Biophys Acta 1583, 237251.Google Scholar
177.Baron, JM, Heise, R, Blaner, WS et al. (2005) Retinoic acid and its 4-oxo metabolites are functionally active in human skin cells in vitro. J Invest Dermatol 125, 143153.Google Scholar
178.Bohm, V & Bitsch, R (1999) Intestinal absorption of lycopene from different matrices and interactions to other carotenoids, the lipid status, and the antioxidant capacity of human plasma. Eur J Nutr 38, 118125.Google Scholar
179.Sass, JO, Masgrau, E, Saurat, JH et al. (1995) Metabolism of oral 9-cis-retinoic acid in the human. Identification of 9-cis-retinoyl-beta-glucuronide and 9-cis-4-oxo-retinoyl-beta-glucuronide as urinary metabolites. Drug Metab Dispos 23, 887891.Google Scholar
180.Barua, AB (1997) Retinoyl beta-glucuronide: a biologically active form of vitamin A. Nutr Rev 55, 259267.Google Scholar
181.Samokyszyn, VM, Gall, WE, Zawada, G et al. (2000) 4-Hydroxyretinoic acid, a novel substrate for human liver microsomal UDP-glucuronosyltransferase(s) and recombinant UGT2B7. J Biol Chem 275, 69086914.Google Scholar
182.Rühl, R, Taner, C, Schweigert, FJ et al. (2010) Serum carotenoids and atopy among children of different ethnic origin living in Germany. Pediatr Allergy Immunol 21, 10721075.Google Scholar
183.Gruber, C, Taner, C, Mihaly, J et al. (2012) Serum retinoic acid and atopy among children of different ethnic origin living in Germany. J Pediatr Gastroenterol Nutr 54, 558560.Google Scholar
184.Rühl, R (2013) Non-pro-vitamin A and pro-vitamin A carotenoids in atopy development. Int Arch Allergy Immunol 161, 99115.Google Scholar
185.Mihaly, J, Marosvolgyi, T, Szegedi, A et al. (2014) Increased FADS2-derived n-6 PUFAs and reduced n-3 PUFAs in plasma of atopic dermatitis patients. Skin Pharmacol Physiol 27, 242248.Google Scholar
186.Mihaly, J, Sonntag, D, Krebiehl, G et al. (2015) Steroid concentrations in patients with atopic dermatitis: reduced plasma dehydroepiandrosterone sulfate and increased cortisone levels. Br J Dermatol 172, 285288.Google Scholar
187.Lucas, R, Mihály, J, Lowe, GM et al. (2018) Reduced carotenoid and retinoid concentrations and altered lycopene isomer ratio in plasma of atopic dermatitis patients. Nutrients 10, 1390.Google Scholar
188.van Vliet, T, van Schaik, F, Schreurs, WH et al. (1996) In vitro measurement of beta-carotene cleavage activity: methodological considerations and the effect of other carotenoids on beta-carotene cleavage. Int J Vitam Nutr Res 66, 7785.Google Scholar
189.Lietz, G, Lange, J & Rimbach, G (2010) Molecular and dietary regulation of beta,beta-carotene 15,15′-monooxygenase 1 (BCMO1). Arch Biochem Biophys 502, 816.Google Scholar
190.Fierce, Y, de Morais Vieira, M, Piantedosi, R et al. (2008) In vitro and in vivo characterization of retinoid synthesis from beta-carotene. Arch Biochem Biophys 472, 126138.Google Scholar
191.Nagao, A (2004) Oxidative conversion of carotenoids to retinoids and other products. J Nutr 134, 237S240S.Google Scholar
192.The Human Protein Atlas (2018) (accessed 20.03.2018 2018).Google Scholar
193.Ford, NA, Moran, NE, Smith, JW et al. (2012) An interaction between carotene-15,15′-monooxygenase expression and consumption of a tomato or lycopene-containing diet impacts serum and testicular testosterone. Int J Cancer 131, E143E148.Google Scholar
194.Boileau, TW, Clinton, SK, Zaripheh, S et al. (2001) Testosterone and food restriction modulate hepatic lycopene isomer concentrations in male F344 rats. J Nutr 131, 17461752.Google Scholar
195.Campbell, JK, Stroud, CK, Nakamura, MT et al. (2006) Serum testosterone is reduced following short-term phytofluene, lycopene, or tomato powder consumption in F344 rats. J Nutr 136, 28132819.Google Scholar
196.Maggio, M, de Vita, F, Lauretani, F et al. (2015) Relationship between carotenoids, retinol, and estradiol levels in older women. Nutrients 7, 65066519.Google Scholar
197.Soderlund, MB, Sjoberg, A, Svard, G et al. (2002) Biological variation of retinoids in man. Scand J Clin Lab Invest 62, 511519.Google Scholar
198.El-Sohemy, A, Baylin, A, Kabagambe, E et al. (2002) Individual carotenoid concentrations in adipose tissue and plasma as biomarkers of dietary intake. Am J Clin Nutr 76, 172179.Google Scholar
199.Tucker, KL, Chen, H, Vogel, S et al. (1999) Carotenoid intakes, assessed by dietary questionnaire, are associated with plasma carotenoid concentrations in an elderly population. J Nutr 129, 438445.Google Scholar
200.van Helden, YG, Godschalk, RW, von Lintig, J et al. (2011) Gene expression response of mouse lung, liver and white adipose tissue to beta-carotene supplementation, knockout of Bcmo1 and sex. Mol Nutr Food Res 55, 14661474.Google Scholar
201.van Helden, YG, Godschalk, RW, Heil, SG et al. (2010) Downregulation of Fzd6 and Cthrc1 and upregulation of olfactory receptors and protocadherins by dietary beta-carotene in lungs of Bcmo1-/- mice. Carcinogenesis 31, 13291337.Google Scholar
202.van Helden, YG, Heil, SG, van Schooten, FJ et al. (2010) Knockout of the Bcmo1 gene results in an inflammatory response in female lung, which is suppressed by dietary beta-carotene. Cell Mol Life Sci 67, 20392056.Google Scholar
203.DiSilvestro, D, Petrosino, J, Aldoori, A et al. (2014) Enzymatic intracrine regulation of white adipose tissue. Horm Mol Biol Clin Investig 19, 3955.Google Scholar
204.Perumal, J, Sriram, S, Lim, HQ et al. (2016) Retinoic acid is abundantly detected in different depots of adipose tissue by SERS. Adipocyte 5, 378383.Google Scholar
205.Bonet, ML, Canas, JA, Ribot, J et al. (2015) Carotenoids and their conversion products in the control of adipocyte function, adiposity and obesity. Arch Biochem Biophys 572, 112125.Google Scholar
206.Amengual, J, Gouranton, E, van Helden, YG et al. (2011) Beta-carotene reduces body adiposity of mice via BCMO1. PLoS ONE 6, e20644.Google Scholar
207.Mercader, J, Ribot, J, Murano, I et al. (2006) Remodeling of white adipose tissue after retinoic acid administration in mice. Endocrinology 147, 53255332.Google Scholar
208.Parvin, SG & Sivakumar, B (2000) Nutritional status affects intestinal carotene cleavage activity and carotene conversion to vitamin A in rats. J Nutr 130, 573577.Google Scholar
209.Paik, J, During, A, Harrison, EH et al. (2001) Expression and characterization of a murine enzyme able to cleave beta-carotene. The formation of retinoids. J Biol Chem 276, 3216032168.Google Scholar
210.van Vliet, T, van Vlissingen, MF, van Schaik, F et al. (1996) Beta-carotene absorption and cleavage in rats is affected by the vitamin A concentration of the diet. J Nutr 126, 499508.Google Scholar
211.Wang, XD, Tang, GW, Fox, JG et al. (1991) Enzymatic conversion of beta-carotene into beta-apo-carotenals and retinoids by human, monkey, ferret, and rat tissues. Arch Biochem Biophys 285, 816.Google Scholar
212.Rubin, LP, Ross, AC, Stephensen, CB et al. (2017) Metabolic effects of inflammation on vitamin A and carotenoids in humans and animal models. Adv Nutr 8, 197212.Google Scholar
213.Rühl, R, Hanel, A, Garcia, AL et al. (2007) Role of vitamin A elimination or supplementation diets during postnatal development on the allergic sensitisation in mice. Mol Nutr Food Res 51, 11731181.Google Scholar
214.Rühl, R, Garcia, A, Schweigert, FJ et al. (2004) Modulation of cytokine production by low and high retinoid diets in ovalbumin-sensitized mice. Int J Vitam Nutr Res 74, 279284.Google Scholar
215.Garcia, AL, Rühl, R, Herz, U et al. (2003) Retinoid- and carotenoid-enriched diets influence the ontogenesis of the immune system in mice. Immunology 110, 180187.Google Scholar
216.Melhus, H, Michaelsson, K, Kindmark, A et al. (1998) Excessive dietary intake of vitamin A is associated with reduced bone mineral density and increased risk for hip fracture. Ann Intern Med 129, 770778.Google Scholar
217.Michaelsson, K, Lithell, H, Vessby, B et al. (2003) Serum retinol levels and the risk of fracture. N Engl J Med 348, 287294.Google Scholar
218.Aage, S, Kiraly, N, Da Costa, K et al. (2015) Neonatal vitamin A supplementation associated with increased atopy in girls. Allergy 70, 985994.Google Scholar
219.Boelsma, E, van de Vijver, LP, Goldbohm, RA et al. (2003) Human skin condition and its associations with nutrient concentrations in serum and diet. Am J Clin Nutr 77, 348355.Google Scholar
220.Yamaguchi, N, Sunto, A, Goda, T et al. (2014) Competitive regulation of human intestinal beta-carotene 15,15′-monooxygenase 1 (BCMO1) gene expression by hepatocyte nuclear factor (HNF)-1alpha and HNF-4alpha. Life Sci 119, 3439.Google Scholar
221.Gong, X, Tsai, SW, Yan, B et al. (2006) Cooperation between MEF2 and PPARgamma in human intestinal beta,beta-carotene 15,15′-monooxygenase gene expression. BMC Mol Biol 7, 7.Google Scholar
222.Gong, X, Marisiddaiah, R & Rubin, LP (2013) beta-carotene regulates expression of beta-carotene 15,15′-monoxygenase in human alveolar epithelial cells. Arch Biochem Biophys 539, 230238.Google Scholar
223.Schupp, M & Lazar, MA (2010) Endogenous ligands for nuclear receptors: digging deeper. J Biol Chem 285, 4040940415.Google Scholar
224.Dhe-Paganon, S, Duda, K, Iwamoto, M et al. (2002) Crystal structure of the HNF4 alpha ligand binding domain in complex with endogenous fatty acid ligand. J Biol Chem 277, 3797337976.Google Scholar
225.Kim, YK, Zuccaro, MV, Costabile, BK et al. (2015) Tissue- and sex-specific effects of beta-carotene 15,15′ oxygenase (BCO1) on retinoid and lipid metabolism in adult and developing mice. Arch Biochem Biophys 572, 1118.Google Scholar
226.Landrier, JF, Kasiri, E, Karkeni, E et al. (2017) Reduced adiponectin expression after high-fat diet is associated with selective up-regulation of ALDH1A1 and further retinoic acid receptor signaling in adipose tissue. FASEB J 31, 203211.Google Scholar
227.Amengual, J, Garcia-Carrizo, FJ, Arreguin, A et al. (2018) Retinoic acid increases fatty acid oxidation and irisin expression in skeletal muscle cells and impacts irisin in vivo. Cell Physiol Biochem 46, 187202.Google Scholar
228.Tourniaire, F, Musinovic, H, Gouranton, E et al. (2015) All-trans retinoic acid induces oxidative phosphorylation and mitochondria biogenesis in adipocytes. J Lipid Res 56, 11001109.Google Scholar
229.Bonet, ML, Ribot, J & Palou, A (2012) Lipid metabolism in mammalian tissues and its control by retinoic acid. Biochim Biophys Acta 1821, 177189.Google Scholar
230.Takeda, K, Sriram, S, Chan, XH et al. (2016) Retinoic acid mediates visceral-specific adipogenic defects of human adipose-derived stem cells. Diabetes 65, 11641178.Google Scholar
231.Brun, PJ, Grijalva, A, Rausch, R et al. (2015) Retinoic acid receptor signaling is required to maintain glucose-stimulated insulin secretion and beta-cell mass. FASEB J 29, 671683.Google Scholar
232.Watzl, B, Bub, A, Brandstetter, BR et al. (1999) Modulation of human T-lymphocyte functions by the consumption of carotenoid-rich vegetables. Br J Nutr 82, 383389.Google Scholar
233.Watzl, B, Bub, A, Blockhaus, M et al. (2000) Prolonged tomato juice consumption has no effect on cell-mediated immunity of well-nourished elderly men and women. J Nutr 130, 17191723.Google Scholar
234.Watzl, B, Bub, A, Briviba, K et al. (2003) Supplementation of a low-carotenoid diet with tomato or carrot juice modulates immune functions in healthy men. Ann Nutr Metab 47, 255261.Google Scholar
235.Borel, P & Desmarchelier, C (2017) Genetic variations associated with vitamin A status and vitamin A bioavailability. Nutrients 9, pii: E246.Google Scholar
236.Ferrucci, L, Perry, JR, Matteini, A et al. (2009) Common variation in the beta-carotene 15,15′-monooxygenase 1 gene affects circulating levels of carotenoids: a genome-wide association study. Am J Hum Genet 84, 123133.Google Scholar
237.Liu, C, Wang, XD, Bronson, RT et al. (2000) Effects of physiological v. pharmacological beta-carotene supplementation on cell proliferation and histopathological changes in the lungs of cigarette smoke-exposed ferrets. Carcinogenesis 21, 22452253.Google Scholar
Figure 0

Fig. 1. (Colour online) Processing of β-carotene during digestion. All factors that impinge on matrix-release, transfer from lipid droplets to mixed micelles, and their diffusion to the enterocyte surface can alter bioaccessibility and thus bioavailability of β-carotene. By contrast, the influence of the colon and its microbiota remains unclear.

Figure 1

Fig. 2. (Colour online) (A) Candidate proteins for β-carotene metabolism within the enterocyte. When genetic variants have been associated with β-carotene bioavailability(25), the encoded proteins are coloured in grey. Dotted lines indicate regulations, i.e. regulation of BCO1 and SR-BI expression by ISX and regulation of chylomicron synthesis by SR-BI and CD36. (B) Candidate proteins that can modulate postprandial blood chylomicron β-carotene concentrations. When genetic variants have been associated with postprandial chylomicron β-carotene response to dietary β-carotene(25), the encoded proteins are coloured in grey. The dotted line indicates that this pathway is assumed but not demonstrated. (C) Proteins involved in the liver metabolism of β-carotene. Note that, to focus on β-carotene and for improved clarity, the fate of chylomicron retinyl esters in the liver is not shown, as well as the liver metabolism of retinol that involves numerous proteins(235). The liver is the hub of β-carotene metabolism: it is the main organ that stores β-carotene and distributes it to the peripheral tissues. β-Carotene reaches the liver mainly as β-carotene and retinyl esters, mainly RP, originating from β-carotene cleavage in the enterocyte and incorporated in chylomicrons. β-Carotene is then mostly stored in hepatic stellate cells. When genetic variants have been associated with blood β-carotene concentrations(80,81,236), the encoded proteins are coloured in grey. βC: β-carotene, ABCA1: ATP binding cassette subfamily A member 1, ABCB1: ATP-binding cassette, sub-family B (MDR/TAP), member 1, ABCG5/G8: ATP-binding cassette, sub-family G member 5 and 8, ATRA: all-trans-retinoic acid, BCO1: β-carotene oxygenase 1, BCO2: β-carotene oxygenase 2, BCO2: β-carotene oxygenase 2, CD36: CD36 molecule, CXCL8: C-X-C motif chemokine ligand 8, ELOVL2: elongation of very long chain fatty acids protein 2, FABP: fatty acid binding protein, GPIHBP1: glycosylphosphatidylinositol-anchored high density lipoprotein binding protein 1, HL: hepatic lipase (encoded by LIPC), HSPGs: heparan sulphate proteoglycans, ISX: intestine specific homoeobox (transcription factor under the control of retinoic acid, regulating expression of SR-BI and BCO1), LDLR: LDL-receptor, LPL: lipoprotein lipase, LRP1: LDL-receptor-related protein 1, MTP: microsomal TAG transfer protein, NPC1L1: Niemann Pick C1-like 1, PKD1L2: polycystin 1-like 2 (gene/pseudogene), RBP4: serum retinol-binding protein, ROL: retinol, RP: retinyl palmitate and other retinyl esters coming from βC cleavage in the enterocyte, RPE65: retinal pigment epithelium-specific 65 kDa protein, SAR1B: secretion associated Ras-related GTPase 1B, SOD2: superoxide dismutase 2, SR-BI: scavenger receptor class B type I, TCF7L2: transcription factor 7-like 2, TTR: transthyretin.

Figure 2

Fig. 3. (Colour online) Metabolism of β-carotene with major metabolites formed in vivo. Involved enzymes, binding proteins, receptors and target genes involved in β-carotene metabolism towards bioactive retinoids. Derivatives marked with ‘*’ have been conclusively identified to be endogenously present. At – all-trans; SCARB1 – scavenger receptor class B type I; CD36 – cluster of density 36; ABCG5 / 8 – ATP binding cassette member 5 / 8; BCO1 – β-carotene oxygenase 1; BCO2 – β-carotene oxygenase 2; LRAT1 / 2 – lecithin retinol acyltransferase, DGAT1 / 2 – diacylglycerol O-acyltransferase 1 / 2; ISX – intestinal transcription factor; STRA6 – stimulated by retinoic acid 6; RBPR2 – retinol-binding protein receptor 2; RDH 5 / 10 – retinol dehydrogenase 5 / 10; DHRS3 / 9 – short-chain dehydrogenase/reductase 3 / 9; RBP1 / 2 / 4 – retinyl-binding protein 1 / 2 / 4; REH – retinyl-esterase; RETSAT – all-trans-retinol 13,14-reductase; ALDH1A1 / 2 / 3 – aldehyde dehydrogenase 1 family, member A1 / 2 / 3; CRABP1 / 2 – cellular-retinoic acid binding protein 1 / 2; RPE65 – retinal pigment epithelium-specific 65 kDa protein; DES1 – sphingolipid delta(4)-desaturase; RLBP1 – retinal-binding protein 1; RAR – retinoic acid receptor; RXR – retinoid-X receptor; TG2 – transglutaminase 2; SCD1 – stearoyl-CoA desaturase / (Δ−9-)desaturase-1; ELOVL6 – elongation of very long chain fatty acids protein 6; HOXB6 / 8 – homoeobox protein 6 / 8, HBEGF – heparin-binding-epidermal growth factor; RARRES2 – retinoic acid receptor responder protein 2 / chemerin; ADIPOQ – adiponectin; UCP1 – uncoupling protein 1, UGT2B7 – UDP-glucuronyltransferase-glucuronosyltransferase-2B7.

Figure 3

Fig. 4. (Colour online) BCO1 localisation and metabolic properties. (A) In vitro kinetic analysis of purified recombinant human BCO1 with β-carotene and β-cryptoxanthin, as published earlier from Lindqvist and Andersson(144). (B) Direct correlation newly calculated based on of serum ATβC to ATRA in children in Germany with different ethical backgrounds(183,184). (C) Direct correlation based on serum ATβC to ATRA levels in Hungarian adults (n 40, Lucas et al.(187)) This figure is just present in the original study in ng/ml, while 1 ng/ml ATRA corresponds to 3.3 nM and 1 ng/ml ATβC to 1.86 nM. (D) Distribution of BCO1 mRNA expression in human tissues, as published previously in Lindqvist and Andersson(144) (PBL – peripheral blood lymphocytes). (E) Differentially expressed genes and pathways by β-carotene v. control diet. Gene expression analysis of different tissues on a control diet supplemented with βC v. control diet (containing adequate vitamin A)-fed mice. A description of the mouse study can be found in van Helden et al.(14). The global transcriptome data were extracted from Gene Expression Omnibus (GEO, Superseries GSE98847), containing lung (GSE98845), liver (GSE98846) and inguinal white adipose tissue (iWAT; GSE27271) and were normalised per tissue and genotype with both sexes included for comparison between sexes. Sex-specific number of differential expressed genes (P < 0·05) are given in number and fold change (FC) of males v. females. (F) ATRA levels in serum (nm) and lung ((pmol/ml / 10−2 m) of control treated (CTRL), low-β-carotene (βC)-diet supplemented (LBC) and high-βC supplemented ferrets (HBC) adapted from Liu et al.(237). Panels A, B, D and F are adapted from van Helden et al.(14) and Lindqvist and Andersson(144) and were permitted to reproduction under copyright.

Figure 4

Fig. 5. Transcriptional regulation of BCO1 metabolism and affected biological processes. Schematic summary of metabolism of the endogenous RAR-activator ATRA starting from ATβC, via all-trans-retinal (ATRAL) to ATRA, which can further activate RAR-RXR-mediated transcriptional signalling. In parallel the newly identified endogenous RXR-ligand 9-cis-13,14-dihydroretinoic acid (9CDHRA) can be created starting from putative carotenoid via putative retinal-analogues to 9CDHRA, which can further activate RXR-hepatocyte nuclear factor (HNF)4α, -PPAR α or -PPARγ-mediated transcriptional signalling. These three receptors (HNF4α, PPARα and PPARγ) can be activated by their ligands, free fatty acids (FFAs) and other metabolites originating from fatty acids. The RAR- or RXR-mediated signalling can positively or negatively alter transcriptional regulated BCO1-expression. LUT, lutein; CAN, canthaxanthin; ZEA, zeaxanthin.

Supplementary material: File

Bohn et al. supplementary material

Table S1

Download Bohn et al. supplementary material(File)
File 94.6 KB