Open access peer-reviewed chapter

Effects of Z-Isomerization on the Bioavailability and Functionality of Carotenoids: A Review

Written By

Masaki Honda, Hayato Maeda, Tetsuya Fukaya and Motonobu Goto

Submitted: 02 March 2018 Reviewed: 04 May 2018 Published: 05 November 2018

DOI: 10.5772/intechopen.78309

From the Edited Volume

Progress in Carotenoid Research

Edited by Leila Queiroz Zepka, Eduardo Jacob-Lopes and Veridiana Vera De Rosso

Chapter metrics overview

1,541 Chapter Downloads

View Full Metrics

Abstract

Carotenoids, the most common fat-soluble plant pigments in nature, are beneficial to human health due to their strong antioxidant activities and abilities to prevent various diseases. Carotenoids have many geometrical isomers forms caused by E/Z-isomerization at arbitrary sites within the multiple conjugated double bonds. Several studies have addressed that the bioavailability as well as the antioxidant, anticancer, and antiatherosclerotic activities of carotenoids varies among the isomers. In addition, those variations differ among carotenoids: Z-isomerization resulted in “positive” or “negative” effect for carotenoids bioavailability and functionality, for example, Z-isomers of lycopene are more bioavailable than the all-E-isomer, whereas the opposite is observed for β-carotene. Thus, to efficiently promote the beneficial effects of carotenoids by ingestion, it is important to have a good understanding of the impact of E/Z-isomerization on the corresponding functional changes. The objective of this contribution is to review the effects of carotenoid Z-isomerization on bioavailability and functionality and describe their differences among carotenoids.

Keywords

  • lycopene
  • β-carotene
  • astaxanthin
  • E/Z-isomer
  • bioavailability
  • antioxidant activity

1. Introduction

Carotenoids are the most common lipid-soluble pigments responsible for the colors of plants, animals, and microorganisms, and over 1100 different types of carotenoids have been characterized so far [1, 2]. Carotenoids can be divided into the following two groups: (1) carotenes, which are nonoxygenated molecules such as lycopene and β-carotene; (2) xanthophylls, which are oxygen-containing molecules such as astaxanthin and fucoxanthin (Figure 1) [3]. The daily consumption of carotenoid-rich foods would be beneficial for human health because of their high antioxidant, anticancer, and antiatherosclerotic activities [4, 5, 6]. Because carotenoids contain numerous conjugated double bonds, many kinds of geometrical isomers are theoretically possible (Figure 1C, E and F). In general, carotenoids in plants occur predominantly in the (all-E)-configuration, whereas the Z-isomers are present in the human body and processed foods in considerable quantity, for example, over 50% of total lycopene is present as the Z-isomers in serum and tissues [7, 8, 9]. Data from several studies have shown that the Z-isomerization of carotenoids induced changes in important properties, such as the bioavailability, antioxidant activity, and anticancer activity [10, 11, 12, 13]. However, these outcomes vary depending on the type of carotenoid: there were cases where the beneficial effects of carotenoids increased or reduced by the Z-isomerization [10, 11, 12, 13, 14, 15]. For example, Z-isomers of lycopene and astaxanthin have higher bioavailability than the all-E-isomers [12, 16], whereas Z-isomers of β-carotene have lower bioavailability than the all-E-isomers [14]. Furthermore, the results may depend on the evaluation method used. For instance, when the antioxidant activity of β-carotene was evaluated based on oxidation of the low-density lipoprotein (LDL), the all-E-isomer showed higher antioxidant activity than the 9Z-isomer [17], whereas the 9Z-isomer showed higher antioxidant activity when evaluated based on antiperoxidative activity [18]. Moreover, the beneficial effects of carotenoids differ between the Z-isomers. For example, when the antioxidant activity of fucoxanthin was evaluated in 1,1-diphenyl-2-picrylhydrazyl (DPPH) radical-scavenging activity assay, the order of activity was 13Z-isomer ≈ 13′Z-isomer > all-E-isomer > 9′Z-isomer [19]. The above findings indicate that a good understanding of the effects of E/Z-isomerization on functional changes is important for increasing the beneficial effects of carotenoid ingestion and for the industrial processing of carotenoids. The objective of this chapter is to highlight the impact of E/Z-isomerization of carotenoids on their bioavailabilities, antioxidant activities, and inhibitory effects against diseases, such as atherogenesis and cancer. Furthermore, aspects of the change factor of the carotenoid bioavailability and functionality, modification of the physicochemical properties of carotenoids by E/Z-isomerization, and Z-isomerization methods used for carotenoids are also discussed in this chapter.

Figure 1.

Chemical structures of (A) (all-E)-lycopene, (B) (all-E)-β-carotene, (C) (all-E)-astaxanthin, (D) (all-E)-fucoxanthin, (E) (9Z)-astaxanthin, and (F) (13Z)-astaxanthin.

Advertisement

2. Effect of Z-isomerization of carotenoids on their bioavailabilities and functionalities

The effects of Z-isomerization on the bioavailability and functionality of eight carotenoids have been investigated thus far, including: (1) the bioavailability and antioxidant activity of lycopene; (2) the antioxidant activity of α-carotene; (3) the bioavailability and antioxidant, antiatherogenenic, and antiatherosclerotic activities of β-carotene; (4) the bioavailability and antioxidant activity of astaxanthin; (5) the antioxidant and pro-apoptotic activities of canthaxanthin; (6) the antioxidant and anticancer activities of fucoxanthin; (7) the bioavailability and antioxidant activity of lutein; and (8) the antioxidant activity of zeaxanthin. The changes caused by Z-isomerization varied according to the parental carotenoid molecules tested and the evaluation method employed. The findings are described in detail below.

2.1. Lycopene

Lycopene is an acyclic carotene (C40H56) that is principally responsible for the bright-red color found abundantly in vegetables and fruits such as tomatoes, guava, and watermelons [3, 9]. Lycopene shows an especially strong antioxidant activity among carotenoids [6] and can significantly reduce the risks for arteriosclerosis, atherogenesis, and many types of cancer (such as prostate and esophageal cancer) [4, 5]. Therefore, in recent years, the use of lycopene in health foods and supplements, and as a natural functional pigment has attracted attention. It is well documented that the bioavailability and antioxidant activity of lycopene are changed by Z-isomerization. Most previous findings have demonstrated that the Z-isomerization of lycopene results in “positive” health effects.

Data from both in vitro and in vivo tests have suggested that Z-isomers of lycopene are more bioavailable than the all-E-isomer. Testing conducted using a diffusion model [20], bile acid micelles [21, 22], human intestinal Caco-2 cells [23], and lymph-cannulated ferrets [21, 22] has provided strong evidence supporting the higher bioavailability of the Z-isomers. Moreover, in humans, the ingestion of foods rich in lycopene Z-isomers resulted in a measurable increase in blood lycopene concentrations compared to a sample abundant in the (all-E)-isomer [12, 24, 25, 26, 27]. For example, Cooperstone et al. [12] investigated the effects of ingesting red tomato juice, which mainly contained (all-E)-lycopene (90% all-E-isomer) and tangerine tomato juice, which mainly contained Z-isomers of lycopene (94% Z-isomers), on plasma lycopene concentrations. Lycopene from the tangerine tomato juice showed approximately 8.5-fold greater bioavailability than lycopene from the red tomato juice. Unlu et al. [25] reported that when comparing two tomato sauces—one rich in all-E-lycopene (95% all-E-isomer) and the other rich in (Z)-lycopene (45% Z-isomers)—that the Z-isomer-rich tomato sauce was approximately 1.5 times more bioavailable than the all-E-isomer-rich sauce. In general, the uptake of carotenoids into intestinal mucosal cells is aided by the formation of bile acid micelles [21, 22, 24, 27]. Thus, it is believed that because lycopene Z-isomers are more soluble in bile acid micelles than the all-E-isomer, they are preferentially incorporated into enterocytes and efficiently form chylomicrons [21, 22]. Indeed, very recently, several reports showed that the solubility of lycopene in oils, organic solvents, and supercritical CO2 (SC-CO2) was significantly improved by Z-isomerization [28, 29, 30, 31, 32]. However, Richelle et al. [33] showed by human oral-dosing tests that the (9Z)- and (13Z)-isomers were less efficiently absorbed than the 5Z- and all-E isomers or were converted into 5Z- and all-E isomers.

Several previous reports have shown that lycopene Z-isomers have higher antioxidant activity than the all-E isomer and that the relative activities of the isomers varied depending on the assay method [10, 11]. Böhm et al. [10] compared the antioxidant activity of (all-E)-lycopene with four unknown Z-isomers by measuring their abilities to reduce radical cations of 2,2′-azinobis(3-ethylbenzothiazoline-6-sulfonic acid) (TEAC assay), and each Z-isomer showed higher antioxidant activity than the all-E-isomer. Müller et al. [11] evaluated the antioxidant activities of (all-E)-, (5Z)-, (9Z)-, (13Z)-, and (7Z,9Z,7′Z,9′Z)-lycopene using four different in vitro assay, namely the TEAC assay, the ferric-reducing antioxidant power (FRAP) assay, the peroxyl radical-scavenging capacity (PSC) assay, and the heme-induced peroxidation of linoleic acid in mildly acidic emulsions mimicking postprandial lipid oxidation in the gastric compartment (MbFeIII-LP) assay. No significant changes were observed among the isomers in the TEAC and FRAP assay. However, the lycopene Z-isomers showed higher antioxidant activities than the all-E isomer in the PSC assay (9Z-isomer > 5Z-isomer ≈ 7Z,9Z,7′Z,9′Z-isomer > 13Z-isomer > all-E-isomer) and in the MbFeIII-LP assay (5Z-isomer > all-E-isomer ≈ 9Z-isomer ≈ 13Z-isomer ≈ 7Z,9Z,7′Z,9′Z-isomer). In TEAC assay, Böhm et al. [10] found that Z-isomers of lycopene had higher antioxidant activity, but Müller et al. [11] concluded that no significant differences occurred among the isomers. These discrepant findings may be explained by the fact that different concentrations of the isomers were used in each study [10, 11].

Based on the above findings, Z-isomerization effectively promotes the beneficial effects of lycopene. Among the Z-isomers, (5Z)-lycopene would have the highest bioavailability [33] and antioxidant activity [11]. Furthermore, the 5Z-isomer has the highest stability of the Z-isomers [34, 35, 36]. Therefore, regarding lycopene, it is very important to increase the 5Z-isomer level and its ingestion. As the Z-isomerization method to increase (5Z)-lycopene efficiently, heating in some alkyl halides [37] and some kinds of oils such as sesame oil [9, 38], light irradiation with photosensitizers [39], and catalytic treatment [20, 40] were effective. Moreover, to our best knowledge, because the effect of ingesting Z-isomer-rich lycopene on inhibiting the development of diseases such as atherogenesis and cancer has not been clarified, further research in that field is expected in the future.

2.2. β-Carotene

β-carotene is a cyclic carotene (C40H56) that is found abundantly in vegetables and fruits, and provides vegetables such as carrots and pumpkins with a deep orange-yellow color [3, 4]. As with other carotenoids, β-carotene has a high antioxidant capacity [6] and preventive effect against various diseases such as cancer and atherogenesis [4, 17]. Furthermore, β-carotene is very important as a retinol precursor, with a high conversion rate [3, 4]. It is also well documented that the bioavailability and antioxidant activity of β-carotene as well as its antiatherogenic activity are changed by Z-isomerization. Most previous studies have shown that the Z-isomerization results in “negative” effect for bioavailability. In contrast, β-carotene Z-isomerization has shown both “positive” and “negative” effects on antioxidant activity, depending on the evaluation method, and “positive” effects have been shown in terms of antiatherogenic activity.

Data from several in vitro and in vivo tests have indicated that Z-isomers of β-carotene are less bioavailable than the all-E-isomer. For example, in vivo tests using Caco-2 cells, HSC-T6 cells, and liver microsomes [14], as well as in vivo tests using ferrets [41] and gerbils [42] have shown this phenomenon. In humans, the intake of Dunaliella salina and Dunaliella bardawil rich in (9Z)-β-carotene showed lower β-carotene bioavailability than foods rich in the all-E-isomer [43, 44, 45, 46, 47, 48]. The effects of the Z-isomer content on the bioavailability were opposite between lycopene and β-carotene. Generally, after carotenoids are extracted from the food matrix and incorporated into mixed micelles, bioaccessible carotenoids can be internalized by enterocytes [21, 22, 24, 27]. The main absorption site of carotenoids is in the duodenum, and several proteins that are temporarily present at the apical membrane mediate selectivity in terms of carotenoid uptake [27, 49, 50, 51]. In vitro experiments with Caco-2 cells showed that carotenoid transport decreased in the following order: β-carotene ≈ α-carotene (50% inhibition) > β-cryptoxanthin ≈ lycopene (20% inhibition) > lutein: zeaxanthin (1:1) (7% inhibition) [49]. Because carotenoid Z-isomers have higher solubility than the all-E-isomers [28, 29, 30, 31, 32], they can incorporate into bile acid micelles more efficiency [21, 22]. Therefore, it is considered that Z-isomers of β-carotene have lower transport efficiency in Caco-2 cell than the all-E-isomers [13]. However, a few studies have suggested that Z-isomers of β-carotene have higher bioavailability than the all-E-isomers, as evaluated using human intestinal Caco-2 cells [52] and ferrets [53]. The use of different delivery systems with the cell model system and animal species might have caused discordant results [15].

Several studies have been conducted to compare the antioxidant activities of (all-E)-β-carotene and the Z-isomers, and the degree of antioxidant activity detected varied according to the assay method. Namely, the 9Z-isomer showed higher antioxidant activity than the all-E-isomer when evaluated in terms of the sensitivity to external oxidants [54], the antiperoxidative activity [18], and oral dose testing in rats [55]. However, the opposite results (or no significant differences) were observed when the antioxidant activities were evaluated by measuring the oxidation of LDL [17] or in TEAC assay [10, 56], and PSC assay [56, 57]. Rodrigues et al. [57] reported that β-carotene Z-isomers were less efficient as peroxyl radical scavengers than the corresponding all-E-isomers: the Z-isomers presented the values about 20% lower than that found for the all-E-isomer, and they addressed that the negative effect may be due to the decreasing of the orbital overlap. Based on the above findings, it is difficult to conclude whether antioxidant activity is enhanced by Z-isomerization of the all-E-isomer.

Moreover, as additional “positive” effects of β-carotene Z-isomers, it has been reported that the 9Z-isomer has higher antiatherogenic activity [58] and antiatherosclerotic activity [59, 60] than the all-E-isomer. On the other hand, there are other “negative” effects. Namely, β-carotene is a very important retinol precursor with a high conversion rate. The (all-E)- and (9Z)-β-carotene can be metabolized respectively to (all-E)-retinoic acid and (9Z)-retinoic acid [61, 62], both of which are active in gene regulation [63, 64]. However, the rates of cleavage of β-carotene isomers to vitamin A and the composition of the respective isomer metabolites vary, that is, (all-E)-β-carotene was the preferred substrate for cleavage to vitamin A when compared with the Z-isomers [61, 65, 66].

Regarding β-carotene, considering that “positive” and “negative” effects are associated with Z-isomerization, it is considered important to use them properly depending on the situation. Besides, as the Z-isomerization method for (all-E)-β-carotene, heating [67, 68], light irradiation with photosensitizers [69], and catalytic treatment [70, 71] were well documented. Moreover, Dunaliella salina and Dunaliella bardawil, which contain a large amount of (9Z)-β-carotene, have been used as Z-isomer-rich materials [43, 44, 45, 46, 47, 48].

2.3. Astaxanthin

Astaxanthin is a xanthophyll (C40H52O4) that is principally responsible for the dark-red color in various microalgae and marine animals [1, 72]. Astaxanthin shows an especially strong antioxidant activity among carotenoids [6] and can significantly reduce the risk of cancer, eye disease, and cardiovascular disease [73, 74]. For instance, astaxanthin protected mice from carcinogenesis of the urinary bladder by reducing the incidence of chemically induced bladder carcinoma and further, astaxanthin supplementation in rats inhibited the stress-induced suppression of tumor-fighting natural killer cells [73]. In addition, astaxanthin is frequently used as an animal and fish feed additive to improve their body colors [75]. Data from several studies have demonstrated that the bioavailability and antioxidant activity of astaxanthin were changed by Z-isomerization.

In terms of the bioavailability, an in vitro test using a simulated digestion model and human intestinal Caco-2 cells [76] and human oral-dosing studies [16, 77] have shown that Z-isomers have higher bioavailability than the all-E-isomer. For example, Yang et al. [76] reported that (13Z)-astaxanthin showed higher bioaccessibility than (9Z)- and (all-E)-astaxanthins using an in vitro-digestion model, and (9Z)-astaxanthin exhibited higher cellular-transport efficiency than (all-E)- and (13Z)-astaxanthin in Caco-2 cell monolayers. However, oral-dosing studies in rainbow trout (Oncorhynchus mykiss) have shown a “negative” effect of astaxanthin Z-isomerization on bioavailability [78, 79]. These results suggest that the bioavailability of carotenoid isomers differs among species. Thus, future studies should seek to establish the biochemical basis for species-specific differences in the utilization of carotenoid isomers.

Although the antioxidant activity measured depends on the assay method employed, many studies have shown “positive” effects. Namely, assay that measure antioxidant enzyme activities, DPPH radical scavenging, oxygen radical-absorption capacity (ORAC), photochemiluminescence (PLC) and peroxidation have shown higher antioxidant activities of astaxanthin Z-isomers than detected for the all-E-isomer [76, 80, 81]. In contrast, when the antioxidant activity was evaluated by a cellular antioxidant activity (CAA) assay, the order of the antioxidant activity was 13Z-isomer > all-E-isomer > 9Z-isomer [81]. The results of these studies suggest that Z-isomers of astaxanthin, especially the 13Z-isomer, have higher antioxidant activity than the all-E-isomer.

Most investigators have concluded that “positive” effects on the bioavailability and antioxidant activity occurred following astaxanthin Z-isomerization. Thus, the ingestion of astaxanthin Z-isomers could be effective in these terms. As with other carotenoids, Z-isomers of astaxanthin could be obtained by heating [81, 82] and catalytic treatment [76, 81, 83] of the all-E-isomer.

2.4. Canthaxanthin

Canthaxanthin is a xanthophyll (C40H52O2) that is principally responsible for the orange-pink color found abundantly in egg yolk and various microbes such as Bradyrhizobium sp. and Halobacterium sp. [84, 85]. Canthaxanthin can significantly reduce the risk of cancer and neurodegenerative disorder [86, 87] and shows strong antioxidant activity [88], that is, canthaxanthin administration decreased mammary tumor volumes in mice [86] and exhibited antiinflammatory activities by increasing the activity of GPX and catalase, thereby reducing the production of IL-1, IL-6, and TNF-α [87]. Furthermore, canthaxanthin is widely used as feed for hens and fish to improve the egg yolk color and the body color, respectively [84, 89]. A few reports have shown the effect of canthaxanthin Z-isomerization on antioxidant activity and functionality. Venugopalan et al. [88] reported that (9Z)-canthaxanthin isolated from Dietzia sp. had higher antioxidant activity, as evaluated by performing DPPH radical-scavenging assay, superoxide radical-scavenging assay and fluorescence assay to detect reactive oxygen species generated in THP-1 cells. Moreover, the (9Z)-isomer exhibited higher pro-apoptotic activity than the all-E-isomer, which was evaluated in THP-1 macrophages [90]. The above literature indicates that Z-isomerization of canthaxanthin has “positive” effects. Canthaxanthin Z-isomerization can be achieved by heating and catalytic treatment [91, 92], and Dietzia sp. can serve as a source of (9Z)-canthaxanthin [88, 89].

2.5. Fucoxanthin

Fucoxanthin is an allenic xanthophyll (C42H58O6) that is found abundantly in edible shellfish and brown seaweeds such as Mactra chinensis and Undaria pinnatifida [1, 93]. Fucoxanthin has high antioxidant capacity [94] and shows anticancer and antiangiogenic activities [95, 96]. For example, fucoxanthin remarkably reduced the viability of human colon cancer cell lines, such as Caco-2, HT-29, and DLD-1 cells [95]. In addition, fucoxanthin has antiobesity and antidiabetic effects [97, 98, 99], for example, administration of Wakame (Undaria pinnatifida) (which is rich in fucoxanthin) significantly suppressed body weight and white adipose tissue weight gain induced by the high fat diet in an obese murine model [98], which has attracted much attention recently in the food industry. The Z-isomerization of (all-E)-fucoxanthin can induce changes in the antioxidant and anticancer activities. Namely, Zhang et al. [19] reported that when the antioxidant activity of fucoxanthin isomers was evaluated by performing DPPH radical-scavenging and superoxide-detection assay, the following relative activities were observed: 13Z-isomer ≈ 13′Z-isomer > all-E-isomer > 9′Z-isomer. Evaluation by performing 2,2-azino-bis(3-ethylbenzothiazoline-6-sulfonic acid) (ABTS) and hydroxyl radical-scavenging assay revealed the following relative activities: 9′Z-isomer > all-E-isomer >13Z-isomer ≈ 13′Z-isomer. Moreover, Kawee-ai et al. [100] showed that when the ratio of the Z-isomer of fucoxanthin increased by 2% from 11 to 13% by heating, the scavenging activities against DPPH, hydrogen peroxide, and superoxide anions, and the reducing power decreased by 21.0, 10.3, 16.0 and 19.7%, respectively. Hence, it is considered that the Z-isomerization of fucoxanthin negatively affects the antioxidant activity. However, Nakazawa et al. [13] demonstrated that Z-isomers of fucoxanthin had higher anticancer activity than the all-E-isomer, as evaluated by measuring the potent inhibitory effects on human promyelocytic leukemia cells (HL-60) and colon cancer cells (Caco-2). To the best of our knowledge, only fucoxanthin was investigated in terms of the effect of Z-isomerization on anticancer activity. Thus, it is expected that such investigation will extend to other carotenoids in the future. Z-isomerization of (all-E)-fucoxanthin has been achieved by heating and light irradiation [100, 101].

2.6. Lutein

Lutein is a xanthophyll (C40H56O2) that is principally responsible for the yellow-orange color found abundantly in vegetables, for example, corn, carrots, kale, and peas, and in egg yolks [102]. Lutein has preventive effects against various diseases such as eye diseases and cardiovascular diseases [102, 103, 104]. In particular, several studies have addressed the role of lutein in reducing the risk of the two most common eye diseases in older people, that is, cataracts and macular degeneration [102, 103, 104]. Only a few reports have shown the effect of lutein Z-isomerization on bioavailability and antioxidant activity [15, 105]. In vitro tests using a digestion model have shown a higher bioaccessibility of Z-isomers of lutein than the all-E-isomer, and a Caco-2 cell monolayer model has shown a lower bioavailability. These results indicated that Z-isomers of lutein are more efficiently incorporated into bile acid micelles, but they have lower transport efficiency in enterocytes via the activities of carotenoid-transport proteins like β-carotene, as described above [15, 27, 49, 50, 51].

In terms of antioxidant activity, the Z-isomers, especially the 13′Z-isomer, have shown higher antioxidant activities than the all-E-isomer in FRAP, DPPH, and ORAC assay, but no significant differences in the activities of the isomers were observed in CAA assay [15]. Since few reports are available regarding the effects of Z-isomerization on lutein bioavailability and functionality, and no such studies have been conducted in humans, further studies are needed to clarify whether Z-isomerization shows “positive” or “negative” effects. Several studies have reported that (all-E)-lutein can be isomerized to the Z-isomers by heating [106, 107] and catalytic treatment [15].

2.7. Other carotenoids

The effects of Z-isomerization on the antioxidant activities of other carotenoids, such as α-carotene and zeaxanthin, were investigated by Böhm et al. [10] by performing TEAC assay. The following relative antioxidant activities of α-carotene stereoisomers were found: 13′Z-isomer > all-E-isomer ≈ 9′Z-isomer > 9Z-isomer ≈ 13Z-isomer, whereas those for zeaxanthin were as follows: all-E-isomer ≈ 13Z-isomer > 9Z-isomer. It is difficult to discern whether Z-isomers of both carotenoids have higher antioxidant activity than the all-E-isomer based on the TEAC assay results alone; thus, further evaluations by multiple testing methods are necessary.

To the best of our knowledge, the effect of Z-isomerization of other important carotenoids such as capsanthin and β-cryptoxanthin (which have large markets and high functionalities) on the bioavailability and functionality has not been reported. Among the over 1100 reported carotenoids found in nature, only the eight carotenoids mentioned above have been characterized in terms of the effects of Z-isomerization, as summarized in Table 1. Thus, further progress in this research area is expected in the future.

CarotenoidEvaluationOverview of resultsEffect*Reference
LycopeneBioavailability/bioaccessibilityZ-Isomers > all-E-isomer, evaluated using a diffusion model+[20]
Z-Isomers > all-E-isomer, evaluated using bile acid micelles and lymph-cannulated ferrets+[21, 22]
Z-Isomers > all-E-isomer, evaluated in Caco-2 cells+[23]
Z-Isomers > all-E-isomer, evaluated in human oral-dosing tests+[12, 24, 25, 26]
Antioxidant activityZ-Isomers > all-E-isomer, evaluated in TEAC assay+[10]
Z-Isomers > all-E-isomer, evaluated in PSC and MbFeIII-LP assay+[11]
All-E-isomer ≈ Z-isomers, evaluated in TEAC and FRAP assay±[11]
α-CaroteneAntioxidant activity13′Z-Isomer > all-E-isomer ≈ 9′Z-isomer > 9Z-isomer ≈ 13Z-isomer, evaluated in TEAC assay±[10]
β-CaroteneBioavailability/bioaccessibilityAll-E-isomer > Z-isomers, evaluated in Caco-2 cells, HSC-T6 cells, and rat liver microsomes[14]
All-E-isomer > Z-isomers, evaluated in ferret oral-dosing test[41]
All-E-isomer > Z-isomers, evaluated in gerbil oral-dosing test[42]
All-E-isomer > 9Z-isomer, evaluated in human oral-dosing tests[43, 44, 45, 46, 47, 48]
9Z-Isomer > all-E-isomer, evaluated in Caco-2 cells+[52]
9Z-Isomer > all-E-isomer, evaluated in the small intestines of ferrets+[53]
Antioxidant activity9Z-Isomer > all-E-isomer, evaluated by measuring the sensitivity to external oxidants+[54]
9Z-Isomer > all-E-isomer, evaluated by determining the antiperoxidative activity+18
9Z-Isomer > all-E-isomer, evaluated in rat oral-dosing tests+[55]
All-E-isomer > 9Z-isomer, evaluated by measuring LDL oxidation[17]
All-E-isomer ≈ 9Z-isomer ≈ 13Z-isomer ≈ 15Z-isomer, evaluated in TEAC assay[10]
All-E-isomer ≈ 9Z-isomer ≈ 13Z-isomer > 15Z-isomer, evaluated in TEAC and PSC assay[56, 57]
Atherogenesis activity9Z-Isomer > all-E-isomer, evaluated in knockout mice+[58]
Atherosclerosis activity9Z-Isomer > all-E-isomer, evaluated in female LDLR−/− and apoE-deficient mice+[59, 60]
AstaxanthinBioavailability/bioaccessibilityZ-Isomers > all-E-isomer, evaluated using a digestion model and Caco-2 cells+[76]
13Z-Isomer > all-E-isomer, 9Z-isomer, evaluated in human oral-dosing test+[77]
Z-Isomers > all-E-isomer, evaluated in human oral-dosing test+[16]
All-E-isomer > Z-isomers, evaluated in rainbow trout (Oncorhynchus mykiss) oral-dosing tests[78, 79]
Antioxidant activityZ-Isomers > all-E-isomer, evaluated in antioxidant enzyme-activity assay+[76]
Z-Isomers > all-E-isomer, evaluated in DPPH and lipid-peroxidation assay+[80]
Z-Isomers > all-E-isomer, evaluated in DPPH, ORAC, and PLC assay+[81]
13Z-Isomer > all-E-isomer >9Z-isomer, evaluated in CAA assay±[81]
CanthaxanthinAntioxidant activity9Z-Isomer > all-E-isomer, evaluated in DPPH, superoxide radical-scavenging, and fluorescence assay+[88]
Pro-apoptotic activity9Z-Isomer > all-E-isomer, evaluated in THP-1 macrophages+[90]
FucoxanthinAntioxidant activity13Z-Isomer ≈ 13′Z-isomer > all-E-isomer > 9′Z-isomer, evaluated in DPPH and superoxide-detection assay±[19]
9′Z-Isomer > all-E-isomer > 13Z-isomer ≈ 13′Z-isomer, evaluated in ABTS and hydroxyl radical-scavenging assay±[19]
Z-Isomers > all-E-isomer, evaluated in DPPH, hydrogen peroxide-scavenging, superoxide anion, and reducing-power assay[100]
Anticancer activityZ-Isomers > all-E-isomer, evaluated in HL-60 cells and Caco-2 cells+[13]
LuteinBioavailability/bioaccessibilityZ-Isomers > all-E-isomer, evaluated using a digestion model+[15]
All-E-isomer > Z-isomers, evaluated in Caco-2 cells[15]
13Z-Isomer > all-E-isomer, evaluated using a digestion model+[105]
Antioxidant activityZ-Isomers > all-E-isomer, evaluated in FRAP assay+[15]
13′Z-Isomer > all-E-isomer ≈ 9Z-isomer, evaluated in DPPH and ORAC assay+[15]
All-E-isomer ≈ Z-isomers, evaluated in CAA assay±[15]
ZeaxanthinAntioxidant activityAll-E-isomer ≈ 13Z-isomer > 9Z-isomers, evaluated in TEAC assay[10]

Table 1.

Summary of the effects of Z-isomerization of different carotenoids on the bioavailability and functionality.

Expected effect of carotenoid Z-isomerization on humans: +, “positive” effect; −, “negative” effect; ±, no change or indetermine.


Advertisement

3. Changes in the physicochemical properties of carotenoids by Z-isomerization

Changes in the bioavailability and functionality of carotenoids after Z-isomerization should have strong correlations with changes in their physicochemical properties. Several reports have shown that the Z-isomerization of carotenoids can induce changes in various properties such as the stability, solubility, and crystallinity. Some computational approaches using a Gaussian program have revealed that the Z-isomerization of carotenoids affected the Gibbs free energy [34, 35, 108, 109], that is, the relative stability of all-E- and mono-Z-isomers were in the following order: all-E-isomer ≈ 5Z-isomer > 9Z-isomer > 13Z-isomer > 15Z-isomer > 7Z-isomer ≈ 11Z-isomer for lycopene [34, 35, 108]; all-E-isomer > 9Z-isomer > 13Z-isomer > 15Z-isomer > 7Z-isomer ≈ 11Z-isome for β-carotene [109]. Thus, (all-E)-carotenoids should be more stable than the Z-isomers, which was confirmed experimentally by Murakami et al. [36] using lycopene isomers. Changes in the Gibbs free energy, stability, of carotenoids following Z-isomerization would affect their antioxidant activities. In addition, there is limited experimental evidence that the Z-isomers of carotenoids such as lycopene, β-carotene, and astaxanthin have higher solubility in vegetable oil, organic solvents and SC-CO2 than the all-E-isomer [28, 29, 30, 31, 32, 110, 111], for example, the solubility of lycopene Z-isomers in ethanol was over 4000 times higher than that of the all-E-isomer [29]. These properties should affect the bioavailability of carotenoids. Namely, Z-isomerization of carotenoids could enhance uptake into bile acid micelles due to an increased solubility; thus, the bioavailability of lycopene and astaxanthin was improved [20, 21, 22]. On the other hand, regarding β-carotene and lutein, whose Z-isomers showed lower bioavailability [15, 41, 42, 43, 44, 45, 46, 47, 48, 105], the uptake into bile acid micelles could potentially be improved by Z-isomerization, but they might have lower transport efficiency in enterocytes due to the activities of several carotenoids transport proteins, which are temporarily present at the apical membrane [27, 49, 50, 51]. In vitro tests of lutein support this hypothesis, that is, the Z-isomers showed higher bioaccessibility than the all-E-isomers in a digestion model [15, 105], whereas the opposite result was obtained in Caco-2 cells [15]. It has been predicted that Z-isomers of lycopene and astaxanthin can be efficiently internalized via carotenoid transporters, based on the results of testing conducted using Caco-2 cells [23, 76]. The abovementioned theory is strongly supported by the observations that, in human blood, over 50% of total lycopene exists in the Z-form, but only 5% of total β-carotene exists in the Z-form [112]. To attain a better understanding of the underlying mechanisms, further study on the uptake process of (Z)-carotenoids in enterocytes by carotenoid transport proteins is necessary. Furthermore, the crystallinity of carotenoids was changed by Z-isomerization: although (all-E)-carotenoids existed in a crystalline state, the Z-isomers were in an amorphous state, which was confirmed by optical observations, differential scanning calorimetry, powder X-ray diffraction, and scanning electron microscopy analyses [20, 28, 29, 113]. The change in crystallinity resulting from Z-isomerization may also influence changes in carotenoid bioavailability and functionality.

Advertisement

4. Conclusions

Z-Isomerization of carotenoids can cause changes in the bioavailability, antioxidant activity, and other functionalities (such as anticancer and antiatherogenic activities), and it may result in “positive” or “negative” effects, which vary according to the type of carotenoid. Although more than 1100 carotenoids are found in nature, only the eight carotenoids discussed above have been investigated in terms of these effects. Thus, further progress in this research area is expected. Furthermore, most investigations have focused on the effects of Z-isomerization of carotenoids on the bioavailability and antioxidant activity, but the Z-isomerization has been shown to enhance the anticancer and antiatherogenic activities of β-carotene and fucoxanthin. Since these data provide important evidence for the roles of carotenoid Z-isomerization in human health, examination of other carotenoids is expected in the future.

References

  1. 1. Maoka T. Carotenoids in marine animals. Marine Drugs. 2011;9:278-293. DOI: 10.3390/md9020278
  2. 2. Yabuzaki J. Carotenoids database: Structures, chemical fingerprints and distribution among organisms. Database. 2017;2017. DOI: 1-11. DOI: 10.1093/database/bax004
  3. 3. Amorim-Carrilho KT, Cepeda A, Fente C, Regal P. Review of methods for analysis of carotenoids. TrAC Trends in Analytical Chemistry. 2014;56:49-73. DOI: 10.1016/j.trac.2013.12.011
  4. 4. Krinsky NI, Johnson EJ. Carotenoid actions and their relation to health and disease. Molecular Aspects of Medicine. 2005;26:459-516. DOI: 10.1016/j.mam.2005.10.001
  5. 5. Xu XR, Zou ZY, Huang YM, Xiao X, Ma L, Lin XM. Serum carotenoids in relation to risk factors for development of atherosclerosis. Clinical Biochemistry. 2012;45:1357-1361. DOI: 10.1016/j.clinbiochem.2012.07.101
  6. 6. Ouchi A, Aizawa K, Iwasaki Y, Inakuma T, Terao J, Nagaoka S, Mukai K. Kinetic study of the quenching reaction of singlet oxygen by carotenoids and food extracts in solution. Development of a singlet oxygen absorption capacity (SOAC) assay method. Journal of Agricultural and Food Chemistry. 2010;58:9967-9978. DOI: 10.1021/jf101947a
  7. 7. Krinsky NI, Russett MD, Handelman GJ, Snodderly DM. Structural and geometrical isomers of carotenoids in human plasma. The Journal of Nutrition. 1990;120:1654-1662. DOI: 10.1093/jn/120.12.1654
  8. 8. Schierle J, Bretzel W, Bühler I, Faccin N, Hess D, Steiner K, Schüep W. Content and isomeric ratio of lycopene in food and human blood plasma. Food Chemistry. 1997;59:459-465. DOI: 10.1016/S0308-8146(96)00177-X
  9. 9. Honda M, Murakami K, Watanabe Y, Higashiura T, Fukaya T, Wahyudiono, Kanda H, Goto M. The E/Z isomer ratio of lycopene in foods and effect of heating with edible oils and fats on isomerization of (all-E)-lycopene. European Journal of Lipid Science and Technology. 2017;119:1600389(1-9). DOI: 10.1002/ejlt.201600389
  10. 10. Böhm V, Puspitasari-Nienaber NL, Ferruzzi MG, Schwartz SJ. Trolox equivalent antioxidant capacity of different geometrical isomers of α-carotene, β-carotene, lycopene, and zeaxanthin. Journal of Agricultural and Food Chemistry. 2002;50:221-226. DOI: 10.1021/jf010888q
  11. 11. Müller L, Goupy P, Fröhlich K, Dangles O, Caris-Veyrat C, Böhm V. Comparative study on antioxidant activity of lycopene (Z)-isomers in different assay. Journal of Agricultural and Food Chemistry. 2011;59:4504-4511. DOI: 10.1021/jf1045969
  12. 12. Cooperstone JL, Ralston RA, Riedl KM, Haufe TC, Schweiggert RM, King SA, Timmers CD, Francis DM, Lesinski GB, Clinton SK, Schwartz SJ. Enhanced bioavailability of lycopene when consumed as cis-isomers from tangerine compared to red tomato juice, a randomized, cross-over clinical trial. Molecular Nutrition & Food Research. 2015;59:658-669. DOI: 10.1002/mnfr.201400658
  13. 13. Nakazawa Y, Sashima T, Hosokawa M, Miyashita K. Comparative evaluation of growth inhibitory effect of stereoisomers of fucoxanthin in human cancer cell lines. Journal of Functional Foods. 2009;1:88-97. DOI: 10.1016/j.jff.2008.09.015
  14. 14. During A, Hussain MM, Morel DW, Harrison EH. Carotenoid uptake and secretion by CaCo-2 cells: β-carotene isomer selectivity and carotenoid interactions. The Journal of Lipid Research. 2002;43:1086-1095. DOI: 10.1194/jlr.M200068-JLR200
  15. 15. Yang C, Fischer M, Kirby C, Liu R, Zhu H, Zhang H, Chen Y, Sun Y, Zhang L, Tsao R. Bioaccessibility, cellular uptake and transport of luteins and assessment of their antioxidant activities. Food Chemistry. 2018;249:66-76. DOI: 10.1016/j.foodchem.2017.12.055
  16. 16. Coral-Hinostroza GN, Ytrestøyl T, Ruyter B, Bjerkeng B. Plasma appearance of unesterified astaxanthin geometrical E/Z and optical R/S isomers in men given single doses of a mixture of optical 3 and 3′R/S isomers of astaxanthin fatty acyl diesters. Comparative Biochemistry and Physiology Part C: Toxicology & Pharmacology. 2004;139:99-110. DOI: 10.1016/j.cca.2004.09.011
  17. 17. Lavy A, Amotz AB, Aviram M. Preferential inhibition of LDL oxidation by the all-trans isomer of β-carotene in comparison with 9-cis β-carotene. Clinical Chemistry and Laboratory Medicine. 1993;31:83-90. DOI: 10.1515/cclm.1993.31.2.83
  18. 18. Levin G, Mokady S. Antioxidant activity of 9-cis compared to all-trans β-carotene in vitro. Free Radical Biology and Medicine. 1994;17:77-82. DOI: 10.1016/0891-5849(94)90009-4
  19. 19. Zhang Y, Fang H, Xie Q, Sun J, Liu R, Hong Z, Yi R, Wu H. Comparative evaluation of the radical-scavenging activities of fucoxanthin and its stereoisomers. Molecules. 2014;19:2100-2113. DOI: 10.3390/molecules19022100
  20. 20. Sun Q, Yang C, Li J, Raza H, Zhang L. Lycopene: Heterogeneous catalytic E/Z isomerization and in vitro bioaccessibility assessment using a diffusion model. Journal of Food Science. 2016;81:C2381-C2389. DOI: 10.1111/1750-3841.13419
  21. 21. Boileau AC, Merchen NR, Wasson K, Atkinson CA, Erdman JW Jr. Cis-lycopene is more bioavailable than trans-lycopene in vitro and in vivo in lymph-cannulated ferrets. The Journal of Nutrition. 1999;129:1176-1181. DOI: 10.1093/jn/129.6.1176
  22. 22. Boileau TWM, Boileau AC, Erdman JW Jr. Bioavailability of all-trans and cis-isomers of lycopene. Experimental Biology and Medicine. 2002;227:914-919. DOI: 10.1177/153537020222701012
  23. 23. Failla ML, Chitchumroonchokchai C, Ishida BK. In vitro micellarization and intestinal cell uptake of cis isomers of lycopene exceed those of all-trans lycopene. The Journal of Nutrition. 2008;138:482-486. DOI: 10.1093/jn/138.3.482
  24. 24. Stahl W, Sies H. Uptake of lycopene and its geometrical isomers is greater from heat-processed than from unprocessed tomato juice in humans. The Journal of Nutrition. 1992;122:2161-2166. DOI: 10.1093/jn/122.11.2161
  25. 25. Unlu NZ, Bohn T, Francis DM, Nagaraja HN, Clinton SK, Schwartz SJ. Lycopene from heat-induced cis-isomer-rich tomato sauce is more bioavailable than from all-trans-rich tomato sauce in human subjects. British Journal of Nutrition. 2007;98(2007):140-146. DOI: 10.1017/S0007114507685201
  26. 26. Burri BJ, Burri BJ, Chapman MH, Neidlinger TR, Seo JS, Ishida BK. Tangerine tomatoes increase total and tetra-cis-lycopene isomer concentrations more than red tomatoes in healthy adult humans. International Journal of Food Sciences and Nutrition. 2009;60:1-16. DOI: 10.1080/09637480701782084
  27. 27. Desmarchelier C, Borel P. Overview of carotenoid bioavailability determinants: From dietary factors to host genetic variations. Trends in Food Science & Technology. 2017;69:270-280. DOI: 10.1016/j.tifs.2017.03.002
  28. 28. Murakami K, Honda M, Wahyudiono, Kanda H, Goto M. Thermal isomerization of (all-E)-lycopene and separation of the Z-isomers by using a low boiling solvent: Dimethyl ether. Separation Science and Technology. 2017;52:2573-2582. DOI: 10.1080/01496395.2017.1374412
  29. 29. Murakami K, Honda M, Takemura R, Fukaya T, Kubota M, Wahyudiono, Kanda H, Goto M. The thermal Z-isomerization-induced change in solubility and physical properties of (all-E)-lycopene. Biochemical and Biophysical Research Communications. 2017;491:317-322. DOI: 10.1016/j.bbrc.2017.07.103
  30. 30. Honda M, Watanabe Y, Murakami K, Takemura R, Fukaya T, Wahyudiono, Kanda H, Goto M. Thermal isomerization pre-treatment to improve lycopene extraction from tomato pulp. LWT-Food Science and Technology. 2017;86:69-75. DOI: 10.1016/j.lwt.2017.07.046
  31. 31. Honda M, Watanabe Y, Murakami K, Hoang NN, Wahyudiono, Kanda H, Goto M. Enhanced lycopene extraction from gac (Momordica cochinchinensis Spreng.) by the Z-isomerization induced with microwave irradiation pre-treatment. European Journal of Lipid Science and Technology. 2018;120:1700293(1-8). DOI: 10.1002/ejlt.201700293
  32. 32. Watanabe Y, Honda M, Higashiura T, Fukaya T, Machmudah S, Wahyudiono , Kanda H, Goto M. Rapid and selective concentration of lycopene Z-isomers from tomato pulp by supercritical CO2 with CO-solvents. Solvent Extraction Research and Development, Japan. 2018;25:47-57
  33. 33. Richelle M, Lambelet P, Rytz A, Tavazzi I, Mermoud AF, Juhel C, Borel P, Bortlik K. The proportion of lycopene isomers in human plasma is modulated by lycopene isomer profile in the meal but not by lycopene preparation. British Journal of Nutrition. 2012;107:1482-1488. DOI: 10.1017/S0007114511004569
  34. 34. Takehara M, Kuwa T, Inoue Y, Kitamura C, Honda M. Isolation and characterization of (15Z)-lycopene thermally generated from a natural source. Biochemical and Biophysical Research Communications. 2015;467:58-62. DOI: 10.1016/j.bbrc.2015.09.122
  35. 35. Honda M, Kudo T, Kuwa T, Higashiura T, Fukaya T, Inoue Y, Kitamura C, Takehara M. Isolation and spectral characterization of thermally generated multi-Z-isomers of lycopene and the theoretically preferred pathway to di-Z-isomers. Bioscience, Biotechnology, and Biochemistry. 2017;81:365-371. DOI: 10.1080/09168451.2016.1249454
  36. 36. Murakami K, Honda M, Takemura R, Fukaya T, Wahyudiono, Kanda H, Goto M. Effect of thermal treatment and light irradiation on the stability of lycopene with high Z-isomers content. Food Chemistry. 2018;250:253-258. DOI: 10.1016/j.foodchem.2018.01.062
  37. 37. Honda M, Takahashi N, Kuwa T, Takehara M, Inoue Y, Kumagai T. Spectral characterisation of Z-isomers of lycopene formed during heat treatment and solvent effects on the E/Z isomerisation process. Food Chemistry. 2015;171:323-329. DOI: 10.1016/j.foodchem.2014.09.004
  38. 38. Honda M, Horiuchi I, Hiramatsu H, Inoue Y, Kitamura C, Fukaya T, Takehara M. Vegetable oil-mediated thermal isomerization of (all-E)-lycopene: Facile and efficient production of Z-isomers. European Journal of Lipid Science and Technology. 2016;118:1588-1592. DOI: 10.1002/ejlt.201500446
  39. 39. Honda M, Igami H, Kawana T, Hayashi K, Takehara M, Inoue Y, Kitamura C. Photosensitized E/Z isomerization of (all-E)-lycopene aiming at practical applications. Journal of Agricultural and Food Chemistry. 2014;62:11353-11356. DOI: 10.1021/jf504502t
  40. 40. Honda M, Kawana T, Takehara M, Inoue Y. Enhanced E/Z isomerization of (all-E)-lycopene by employing iron(III) chloride as a catalyst. Journal of Food Science. 2015;80:C1453-C1459. DOI: 10.1111/1750-3841.12916
  41. 41. Erdman JW, Thatcher AJ, Hofmann NE, Lederman JD, Block SS, Lee CM, Mokady S. All-trans β-carotene is absorbed preferentially to 9-cis β-carotene, but the latter accumulates in the tissues of domestic ferrets (Mustela putorius puro). The Journal of Nutrition. 1998;128:2009-2013. DOI: 10.1093/jn/128.11.2009
  42. 42. Deming DM, Teixeira SR, Erdman JW. All-trans β-carotene appears to be more bioavailable than 9-cis or 13-cis β-carotene in gerbils given single oral doses of each isomer. The Journal of Nutrition. 2002;132:2700-2708. DOI: 10.1093/jn/132.9.2700
  43. 43. Stahl W, Schwarz W, Sies H. Human serum concentrations of all-trans β-and α-carotene but not 9-cis β-carotene increase upon ingestion of a natural isomer mixture obtained from Dunaliella salina (Betatene). The Journal of Nutrition. 1993;123:847-851. DOI: 10.1093/jn/123.5.847
  44. 44. Stahl W, Schwarz W, von Laar J, Sies H. All-trans β-carotene preferentially accumulates in human chylomicrons and very low density lipoproteins compared with the 9-cis geometrical isomer. The Journal of Nutrition. 1995;125:2128-2133. DOI: 10.1093/jn/125.8.2128
  45. 45. Gaziano JM, Johnson EJ, Russell RM, Manson JE, Stampfer MJ, Ridker PM, Frei B, Hennekens CM, Krinsky NI. Discrimination in absorption or transport of β-carotene isomers after oral supplementation with either all-trans-or 9-cis-β-carotene. The American Journal of Clinical Nutrition. 1995;61:1248-1252. DOI: 10.1093/ajcn/61.6.1248
  46. 46. Johnson EJ, Qin J, Krinsky NI, Russell RM. β-Carotene isomers in human serum, breast milk and buccal mucosa cells after continuous oral doses of all-trans and 9-cis β-carotene. The Journal of Nutrition. 1997;127:1993-1999. DOI: 10.1093/jn/127.10.1993
  47. 47. Johnson EJ, Krinsky NI, Russell RM. Serum response of all-trans and 9-cis isomers of β-carotene in humans. Journal of the American College of Nutrition. 1996;15:620-624. DOI: 10.1080/07315724.1996.10718639
  48. 48. Tamai H, Morinobu T, Murata T, Manago M, Mino M. 9-cis β-carotene in human plasma and blood cells after ingestion of β-carotene. Lipids. 1995;30:493-498. DOI: 10.1007/BF02537022
  49. 49. During A, Dawson HD, Harrison EH. Carotenoid transport is decreased and expression of the lipid transporters SR-BI, NPC1L1, and ABCA1 is downregulated in Caco-2 cells treated with ezetimibe. The Journal of Nutrition. 2005;135:2305-2312. DOI: 10.1093/jn/135.10.2305
  50. 50. Kotake-Nara E. Intestinal absorption of carotenoid. Journal of Lipid Nutrition. 2012;21:35-43. DOI: 10.4010/jln.21.35
  51. 51. Yonekura L, Nagao A. Intestinal absorption of dietary carotenoids. Molecular Nutrition & Food Research. 2007;51:107-115. DOI: 10.1002/mnfr.200600145
  52. 52. Ferruzzi MG, Lumpkin JL, Schwartz SJ, Failla M. Digestive stability, micellarization, and uptake of β-carotene isomers by Caco-2 human intestinal cells. Journal of Agricultural and Food Chemistry. 2006;54:2780-2785. DOI: 10.1021/jf0530603
  53. 53. Hébuterne X, Wang XD, Johnson EJ, Krinsky NI, Russell RM. Intestinal absorption and metabolism of 9-cis-β-carotene in vivo: Biosynthesis of 9-cis-retinoic acid. The Journal of Lipid Research. 1995;36:1264-1273
  54. 54. Jiménez C, Pick U. Differential reactivity of β-carotene isomers from Dunaliella bardawil toward oxygen radicals. Plant Physiology. 1993;101:385-390. DOI: 10.1104/pp.101.2.385
  55. 55. Levin G, Yeshurun M, Mokady S. In vivo antiperoxidative effect of 9-cis β-carotene compared with that of the all-trans isomer. Nutrition and Cancer. 1997;27:293-297. DOI: 10.1080/01635589709514540
  56. 56. Mueller L, Boehm V. Antioxidant activity of β-carotene compounds in different in vitro assay. Molecules. 2011;16:1055-1069. DOI: 10.3390/molecules16021055
  57. 57. Rodrigues E, Mariutti LR, Chisté RC, Mercadante AZ. Development of a novel micro-assay for evaluation of peroxyl radical scavenger capacity: Application to carotenoids and structure–activity relationship. Food Chemistry. 2012;135:2103-2111. DOI: 10.1016/j.foodchem.2012.06.074
  58. 58. Harari A, Harats D, Marko D, Cohen H, Barshack I, Kamari Y, Gonen A, Gerber Y, Ben-Amotz A, Shaish A. A 9-cis β-carotene–enriched diet inhibits atherogenesis and fatty liver formation in LDL receptor knockout mice. The Journal of Nutrition. 2008;138:1923-1930. DOI: 10.1093/jn/138.10.1923
  59. 59. Relevy NZ, Rühl R, Harari A, Grosskopf I, Barshack I, Ben-Amotz A, Nir U, Gottlieb H, Kamari Y, Harats D, Shaish A. 9-cis β-Carotene inhibits atherosclerosis development in female LDLR−/− mice. Functional Foods in Health & Disease. 2015;5:67-79
  60. 60. Harari A, Abecassis R, Relevi N, Levi Z, Ben-Amotz A, Kamari Y, Harats A, Shaish, A. Prevention of atherosclerosis progression by 9-cis-β-carotene rich alga Dunaliella in apoE-deficient mice. BioMed Research International. 2013;2013:169517(1-7). DOI: 10.1155/2013/169517
  61. 61. Nagao A, Olson JA. Enzymatic formation of 9-cis, 13-cis, and all-trans retinals from isomers of β-carotene. FASEB Journal. 1994;8:968-973. DOI: 10.1096/fasebj.8.12.8088462
  62. 62. Wang XD, Krinsky NI, Benotti PN, Rusell RM. Biosynthesis of 9-cis-retinoic acid from 9-cis-β-carotene in human intestinal mucosa in vitro. Archives of Biochemistry and Biophysics. 1994;313:150-155. DOI: 10.1006/abbi.1994.1371
  63. 63. Heyman RA, Mangelsdorf DJ, Dyck JA, Stein RB, Eichele G, Evans RM, Thaller C. 9-Cis retinoic acid is a high affinity ligand for the retinoid X receptor. Cell. 1992;68:397-406. DOI: 10.1016/0092-8674(92)90479-V
  64. 64. Levin AA, Sturzenbecker LJ, Kazmer S, Bosakowski T, Huselton C, Allenby G, Speck J, Kratzeisen CL, Rosenberger M, Lovey E, Grippo JF. 9-Cis retinoic acid stereoisomer binds and activates the nuclear receptor RXRα. Nature. 1992;355:359-361. DOI: 10.1038/355359a0
  65. 65. Minguez-Mosquera MI, Hornero-Mendez D, Perez-Galvez A. Carotenoids and provitamin A in functional foods. In: Hurst WJ, editor. Methods of Analysis for Functional Foods and Nutraceuticals. Oxford: Taylor & Francis; 2002. pp. 101-157. DOI: 10.1201/9781420014679.ch3
  66. 66. Schieber A, Carle R. Occurrence of carotenoid cis-isomers in food: Technological, analytical, and nutritional implications. Trends in Food Science & Technology. 2005;16:416-422. DOI: 10.1016/j.tifs.2005.03.018
  67. 67. Lemmens L, De Vleeschouwer K, Moelants KR, Colle IJ, Van Loey AM, Hendrickx ME. β-Carotene isomerization kinetics during thermal treatments of carrot puree. Journal of Agricultural and Food Chemistry. 2010;58:6816-6824. DOI: 10.1021/jf100449t
  68. 68. Knockaert G, Pulissery SK, Lemmens L, Van Buggenhout S, Hendrickx M, Van Loey A. Carrot β-carotene degradation and isomerization kinetics during thermal processing in the presence of oil. Journal of Agricultural and Food Chemistry. 2012;60:10312-10319. DOI: 10.1021/jf3025776
  69. 69. Kuki M, Koyama Y, Nagae H. Triplet-sensitized and thermal isomerization of all-trans, 7-cis, 9-cis, 13-cis and 15-cis isomers of β-carotene: Configurational dependence of the quantum yield of isomerization via the T1 state. The Journal of Physical Chemistry. 1991;95:7171-7180. DOI: 10.1021/j100172a016
  70. 70. Gao Y, Kispert LD. Reaction of carotenoids and ferric chloride: Equilibria, isomerization, and products. The Journal of Physical Chemistry B. 2003;107:5333-5338. DOI: 10.1021/jp034063q
  71. 71. Gao Y, Kispert LD, Konovalova TA, Lawrence JN. Isomerization of carotenoids in the presence of MCM-41 molecular sieves: EPR and HPLC studies. The Journal of Physical Chemistry B. 2004;108:9456-9462. DOI: 10.1021/jp036091e
  72. 72. Yuan JP, Peng J, Yin K, Wang JH. Potential health-promoting effects of astaxanthin: A high-value carotenoid mostly from microalgae. Molecular Nutrition & Food Research. 2011;55:150-165. DOI: 10.1002/mnfr.201000414
  73. 73. Guerin M, Huntley ME, Olaizola M. Haematococcus astaxanthin: Applications for human health and nutrition. Trends in Biotechnology. 2003;21:210-216. DOI: 10.1016/S0167-7799(03)00078-7
  74. 74. Pashkow FJ, Watumull DG, Campbell CL. Astaxanthin: A novel potential treatment for oxidative stress and inflammation in cardiovascular disease. American Journal of Cardiology. 2008;101:S58-S68. DOI: 10.1016/j.amjcard.2008.02.010
  75. 75. Ambati RR, Phang SM, Ravi S, Aswathanarayana RG. Astaxanthin: Sources, extraction, stability, biological activities and its commercial applications—A review. Marine Drugs. 2014;12:128-152. DOI: 10.3390/md12010128
  76. 76. Yang C, Zhang H, Liu R, Zhu H, Zhang L, Tsao R. Bioaccessibility, cellular uptake, and transport of astaxanthin isomers and their antioxidative effects in human intestinal epithelial Caco-2 cells. Journal of Agricultural and Food Chemistry. 2017;65:10223-10232. DOI: 10.1021/acs.jafc.7b04254
  77. 77. Østerlie M, Bjerkeng B, Liaaen-Jensen S. Plasma appearance and distribution of astaxanthin E/Z and R/S isomers in plasma lipoproteins of men after single dose administration of astaxanthin. The Journal of Nutritional Biochemistry. 2000;11:482-490. DOI: 10.1016/S0955-2863(00)00104-2
  78. 78. Bjerkeng B, Følling M, Lagocki S, Storebakken T, Olli JJ, Alsted N. Bioavailability of all-E-astaxanthin and Z-isomers of astaxanthin in rainbow trout (Oncorhynchus mykiss). Aquaculture. 1997;157:63-82. DOI: 10.1016/S0044-8486(97)00146-4
  79. 79. Østerlie M, Bjerkeng B, Liaaen-Jensen S. Accumulation of astaxanthin all-E, 9Z and 13Z geometrical isomers and 3 and 3′ RS optical isomers in rainbow trout (Oncorhynchus mykiss) is selective. The Journal of Nutrition. 1999;129:391-398. DOI: 10.1093/jn/129.2.391
  80. 80. Liu X, Osawa T. Cis astaxanthin and especially 9-cis astaxanthin exhibits a higher antioxidant activity in vitro compared to the all-trans isomer. Biochemical and Biophysical Research Communications. 2007;357:187-193. DOI: 10.1016/j.bbrc.2007.03.120
  81. 81. Yang C, Zhang L, Zhang H, Sun Q, Liu R, Li J, Wu L, Tsao R. Rapid and efficient conversion of all-E-astaxanthin to 9Z- and 13Z-isomers and assessment of their stability and antioxidant activities. Journal of Agricultural and Food Chemistry. 2017;65:818-826. DOI: 10.1021/acs.jafc.6b04962
  82. 82. Yuan JP, Chen F. Kinetics for the reversible isomerization reaction of trans-astaxanthin. Food Chemistry. 2011;73:131-137. DOI: 10.1016/S0308-8146(01)00107-8
  83. 83. Zhao L, Chen F, Zhao G, Wang Z, Liao X, Hu X. Isomerization of trans-astaxanthin induced by copper (II) ion in ethanol. Journal of Agricultural and Food Chemistry. 2005;53:9620-9623. DOI: 10.1021/jf0517750
  84. 84. Cho JH, Zhang ZF, Kim IH. Effects of canthaxanthin on egg production, egg quality, and egg yolk color in laying hens. Journal of Agricultural Science. 2012;5:269-274. DOI: 10.5539/jas.v5n1p269
  85. 85. Malik K, Tokkas J, Goyal S. Microbial pigments: A review. International Journal of Microbial Resource Technology. 2012;1:361-365
  86. 86. Chew BP, Park JS, Wong MW, Wong TS. A comparison of the anticancer activities of dietary β-carotene, canthaxanthin and astaxanthin in mice in vivo. Anticancer Research. 1999;19:1849-1853
  87. 87. Chan KC, Mong MC, Yin MC. Antioxidative and anti-inflammatory neuroprotective effects of astaxanthin and canthaxanthin in nerve growth factor differentiated PC12 cells. Journal of Food Science. 2009;74:H225-H231. DOI: 10.1111/j.1750-3841.2009.01274.x
  88. 88. Venugopalan V, Tripathi SK, Nahar P, Saradhi PP, Das RH, Gautam HK. Characterization of canthaxanthin isomers isolated from a new soil Dietzia sp. and their antioxidant activities. Journal of Microbiology and Biotechnology. 2013;23:237-245. DOI: 10.4014/jmb.1203.03032
  89. 89. Li M, Rahman MM, Wu B, Lin YC. Effects of dietary canthaxanthin on growth and body colour of blood parrot cichlid Amphilophus citrinellus × Paraneetroplus synspilus. Aquaculture International. 2017;25:705-713. DOI: 10.1007/s10499-016-0068-z
  90. 90. Venugopalan V, Verma N, Gautam HK, Saradhi PP, Das RH. 9-cis-Canthaxanthin exhibits higher pro-apoptotic activity than all-trans-canthaxanthin isomer in THP-1 macrophage cells. Free Radical Research. 2009;43:100-105. DOI: 10.1080/10715760802616668
  91. 91. Qiu D, Zhu WL, Tang CK, Shi LF, Gao HQ. Identification of the composition of isomeric canthaxanthin sample by NMR, HPLC, and mass spectrometry. Food Analytical Methods. 2014;7:597-605. DOI: 10.1007/s12161-013-9660-2
  92. 92. Sundquist AR, Hanusch M, Stahl W, Sies H. Cis/trans isomerization of carotenoids by the triplet carbonyl source 3-hydroxymethyl-3,4,4-trimethyl-1,2-dioxetane. Photochemistry and Photobiology. 1993;57:785-791. DOI: 10.1111/j.1751-1097.1993.tb09211.x
  93. 93. Maeda H, Hosokawa M, Sashima T, Funayama K, Miyashita K. Fucoxanthin from edible seaweed, Undaria pinnatifida, shows antiobesity effect through UCP1 expression in white adipose tissues. Biochemical and Biophysical Research Communications. 2005;332:392-397. DOI: 10.1016/j.bbrc.2005.05.002
  94. 94. Sachindra NM, Sato E, Maeda H, Hosokawa M, Niwano Y, Kohno M, Miyashita K. Radical scavenging and singlet oxygen quenching activity of marine carotenoid fucoxanthin and its metabolites. Journal of Agricultural and Food Chemistry. 2007;55:8516-8522. DOI: 10.1021/jf071848a
  95. 95. Hosokawa M, Kudo M, Maeda H, Kohno H, Tanaka T, Miyashita K. Fucoxanthin induces apoptosis and enhances the antiproliferative effect of the PPARγ ligand, troglitazone, on colon cancer cells. Biochimica et Biophysica Acta (BBA)—General Subject. 2004;1675:113-119. DOI: 10.1016/j.bbagen.2004.08.012
  96. 96. Sugawara T, Matsubara K, Akagi R, Mori M, Hirata T. Antiangiogenic activity of brown algae fucoxanthin and its deacetylated product, fucoxanthinol. Journal of Agricultural and Food Chemistry. 2006;54:9805-9810. DOI: 10.1021/jf062204q
  97. 97. Maeda H. Nutraceutical effects of fucoxanthin for obesity and diabetes therapy: A review. Journal of Oleo Science. 2015;64:125-132. DOI: 10.5650/jos.ess14226
  98. 98. Maeda H, Hosokawa M, Sashima T, Murakami-Funayama K, Miyashita K. Anti-obesity and anti-diabetic effects of fucoxanthin on diet-induced obesity conditions in a murine model. Molecular Medicine Reports. 2009;2:897-902. DOI: 10.3892/mmr_00000189
  99. 99. Maeda H, Kanno S, Kodate M, Hosokawa M, Miyashita K. Fucoxanthinol, metabolite of fucoxanthin, improves obesity-induced inflammation in adipocyte cells. Marine Drugs. 2015;13:4799-4813. DOI: 10.3390/md13084799
  100. 100. Kawee-ai A, Kuntiya A, Kim SM. Anticholinesterase and antioxidant activities of fucoxanthin purified from the microalga Phaeodactylum tricornutum. Natural Product Communications. 2013;8:1381-1386
  101. 101. Zhao D, Kim SM, Pan CH, Chung D. Effects of heating, aerial exposure and illumination on stability of fucoxanthin in canola oil. Food Chemistry. 2014;145:505-513. DOI: 10.1016/j.foodchem.2013.08.045
  102. 102. Abdel-Aal ESM, Akhtar H, Zaheer K, Ali R. Dietary sources of lutein and zeaxanthin carotenoids and their role in eye health. Nutrients. 2013;5:1169-1185. DOI: 10.3390/nu5041169
  103. 103. Mares-Perlman JA, Millen AE, Ficek TL, Hankinson SE. The body of evidence to support a protective role for lutein and zeaxanthin in delaying chronic disease. Overview. The Journal of Nutrition. 2002;132:518S-524S. DOI: 10.1093/jn/132.3.518S
  104. 104. Mozaffarieh M, Sacu S, Wedrich A. The role of the carotenoids, lutein and zeaxanthin, in protecting against age-related macular degeneration: A review based on controversial evidence. Nutrition Journal. 2003;2:1-8. DOI: 10.1186/1475-2891-2-20
  105. 105. Rodrigues DB, Mariutti LRB, Mercadante AZ. An in vitro digestion method adapted for carotenoids and carotenoid esters: Moving forward towards standardization. Food & Function. 2016;7:4992-5001. DOI: 10.1039/C6FO01293K
  106. 106. Subagio A, Morita N, Sawada S. Thermal isomerization of all-trans-lutein in a benzene solution. Bioscience, Biotechnology, and Biochemistry. 1998;62:2453-2456. DOI: 10.1271/bbb.62.2453
  107. 107. Updike AA, Schwartz SJ. Thermal processing of vegetables increases cis isomers of lutein and zeaxanthin. Journal of Agricultural and Food Chemistry. 2003;51:6184-6190. DOI: 10.1021/jf030350f
  108. 108. Chasse GA, Mak ML, Deretey E, Farkas I, Torday LL, Papp JG, Sarma DSR, Agarwal A, Chakravarthi S, Agarwal S, Rao AV. An ab initio computational study on selected lycopene isomers. Journal of Molecular Structure: THEOCHEM. 2001;571:27-37. DOI: 10.1016/S0166-1280(01)00424-9
  109. 109. Guo WH, Tu CY, Hu CH. Cis-trans isomerizations of β-carotene and lycopene: A theoretical study. The Journal of Physical Chemistry B. 2008;112:12158-12167. DOI: 10.1021/jp8019705
  110. 110. Gamlieli-Bonshtein I, Korin E, Cohen S. Selective separation of cis-trans geometrical isomers of β-carotene via CO2 supercritical fluid extraction. Biotechnology and Bioengineering. 2002;80:169-174. DOI: 10.1002/bit.10357
  111. 111. Kaga K, Honda M, Adachi T, Honjo M, Wahyudiono, Kanda H, Goto M. Nanoparticle formation of PVP/astaxanthin inclusion complex by solution-enhanced dispersion by supercritical fluids (SEDS): Effect of PVP and astaxanthin Z-isomer content. The Journal of Supercritical Fluids. 2018;36:44-51. DOI: 10.1016/j.supflu.2018.02.008
  112. 112. Stahl W, Schwarz W, Sundquist AR, Sies H. cis-trans Isomers of lycopene and β-carotene in human serum and tissues. Archives of Biochemistry and Biophysics. 1992;294:173-177. DOI: 10.1016/0003-9861(92)90153-N
  113. 113. Hempel J, Schädle CN, Leptihn S, Carle R, Schweiggert RM. Structure related aggregation behavior of carotenoids and carotenoid esters. Journal of Photochemistry and Photobiology A: Chemistry. 2016;317:161-174. DOI: 10.1016/j.jphotochem.2015.10.024

Written By

Masaki Honda, Hayato Maeda, Tetsuya Fukaya and Motonobu Goto

Submitted: 02 March 2018 Reviewed: 04 May 2018 Published: 05 November 2018