Next Article in Journal
Analyzing the Association between Student Weight Status and School Meal Participation: Evidence from the School Nutrition and Meal Cost Study
Next Article in Special Issue
The Effectiveness of Smoking Cessation, Alcohol Reduction, Diet and Physical Activity Interventions in Improving Maternal and Infant Health Outcomes: A Systematic Review of Meta-Analyses
Previous Article in Journal
Caloric Intake in Renal Patients: Repercussions on Mineral Metabolism
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Impact of Maternal Obesity on the Metabolism and Bioavailability of Polyunsaturated Fatty Acids during Pregnancy and Breastfeeding

by
Daniela Álvarez
1,
Yasna Muñoz
1,
Macarena Ortiz
1,
Manuel Maliqueo
1,
Raphaël Chouinard-Watkins
2 and
Rodrigo Valenzuela
2,3,*
1
Endocrinology and Metabolism Laboratory, West Division, Faculty of Medicine, University of Chile, Santiago 8380000, Chile
2
Department of Nutritional Sciences, Faculty of Medicine, University of Toronto, Toronto, ON M5S1A8, Canada
3
Nutrition Department, Faculty of Medicine, University of Chile, Santiago 8380000, Chile
*
Author to whom correspondence should be addressed.
Nutrients 2021, 13(1), 19; https://doi.org/10.3390/nu13010019
Submission received: 4 November 2020 / Revised: 9 December 2020 / Accepted: 16 December 2020 / Published: 23 December 2020
(This article belongs to the Special Issue Maternal Obesity and Nutrients)

Abstract

:
Prenatal and postnatal development are closely related to healthy maternal conditions that allow for the provision of all nutritional requirements to the offspring. In this regard, an appropriate supply of fatty acids (FA), mainly n-3 and n-6 long-chain polyunsaturated fatty acids (LCPUFA), is crucial to ensure a normal development, because they are an integral part of cell membranes and participate in the synthesis of bioactive molecules that regulate multiple signaling pathways. On the other hand, maternal obesity and excessive gestational weight gain affect FA supply to the fetus and neonate, altering placental nutrient transfer, as well as the production and composition of breast milk during lactation. In this regard, maternal obesity modifies FA profile, resulting in low n-3 and elevated n-6 PUFA levels in maternal and fetal circulation during pregnancy, as well as in breast milk during lactation. These modifications are associated with a pro-inflammatory state and oxidative stress with short and long-term consequences in different organs of the fetus and neonate, including in the liver, brain, skeletal muscle, and adipose tissue. Altogether, these changes confer to the offspring a higher risk of developing obesity and its complications, as well as neuropsychiatric disorders, asthma, and cancer. Considering the consequences of an abnormal FA supply to offspring induced by maternal obesity, we aimed to review the effects of obesity on the metabolism and bioavailability of FA during pregnancy and breastfeeding, with an emphasis on LCPUFA homeostasis.

1. Introduction

Maternal obesity and excessive gestational weight gain (GWG) are frequently associated with alterations in carbohydrate and lipid metabolism, abnormal levels of pregnancy hormones, and a pro-inflammatory state. These disturbances increase the risk of maternal and fetal complications such as gestational diabetes mellitus, hypertensive disorders of pregnancy, cesarean delivery, lung disease, miscarriage, still birth, fetal chromosomic anomalies, preterm birth, and fetal macrosomia [1,2]. Moreover, adiposity excess alters the placental nutrient transfer and modifies the composition of breast milk, affecting the development and genetic programing of multiple fetal organs (liver, adipose tissue, skeletal muscle, and brain, among others) (Figure 1). Altogether, these changes confer to the offspring a higher risk of obesity and related complications, as well as of neuropsychiatric disorders, asthma, and cancer [3,4,5,6,7].
During pregnancy and lactation, all the nutritional requirements necessary to ensure an appropriate growth of the fetus and neonate are transferred from the mother [8]. Fetal development is intimately linked to the placental exchange function, such that newborn size reflects the net nutrient transfer through the placenta [9]. Likewise, breast milk provides fats, proteins, carbohydrates, immune cells, and bioactive molecules for optimal infant growth [10,11,12].
Fatty acids (FA) participate in fat mass accretion, regulation of metabolic pathways, and visual and brain development. Depending on the number of double bounds, FA can be classified as saturated (SFA), monounsaturated (MUFA), or polyunsaturated (PUFA) [13]. Of these, n-3 and n-6 long chain PUFA (LCPUFA) are of particular interest because they are an integral part of cell membranes and participate in the synthesis of bioactive molecules that regulate multiple signaling pathways [13]. The supply of essential FA to the fetus will depend entirely on maternal consumption, placental transport and metabolism of FA [14]. Therefore, a balanced diet for pregnant and lactating women must include an appropriate supply of FA. In humans, the synthesis of n-3 and n-6 PUFA requires α–linolenic acid (C18:3n-3, ALA) and linoleic acid (C18:2n-6, LA), respectively. These FA are considered essential and cannot be synthesized in mammals due to a lack of the enzyme delta-12 desaturase (FADS2) and delta-15 desaturase (FADS3) [15]. ALA and LA are desaturated by desaturases (FADS1 or Δ-5 and FADS2 or Δ-6) and elongated by elongases (ELOVL 2 and 5) to produce, notably, docosahexaenoic acid (C22:6n-3, DHA) or arachidonic acid (C20:4n-6, AA). AA and DHA are crucial for the formation and maturation of the brain and other organs during fetal and neonatal development. These characteristics emphasize the importance of an adequate intake of LCPUFA during these periods. Some of the LCPUFA elongation products can also be oxidized through peroxisomal β-oxidation [16].
In general, obesity is characterized by high energy intake due to excessive consumption of food rich in refined carbohydrates and saturated fats, and low consumption of sea foods, vegetables, fruits, and legumes (important sources of dietary fiber and natural antioxidants) [17,18,19]. This dietary pattern is associated with an imbalance in FA intake which is reflected by an excessive consumption of n-6 PUFA and low n-3 PUFA consumption [19]. In this regard, the main sources of DHA are fatty fish and seaweed [20]. Moreover, measurements using traceable carbon 13 (13C) FA stable isotopes and the quantification of the gene and protein expression of FA transporters in the placenta indicate that obesity affects the transplacental transport of FA that are necessary for fetal development, among which DHA is one of the most important [21]. Breast milk adjusts to neonatal demand by varying the composition of bioactive compounds and nutrients, such as FA, during lactation [22].
The short- and long-term consequences of maternal obesity are important public health issues. For this reason, extensive research around the world has been focused on the role of LCPUFA in fetal and neonatal growth, with the purpose of improving nutritional quality during these periods. In view of these considerations, the aim of this review was to summarize the available information about the impact of obesity on the metabolism and bioavailability of PUFA during pregnancy and breastfeeding.

2. Material and Methods

This review included both in vivo and in vitro studies aimed at evaluating the effect of obesity on the metabolism and bioavailability of PUFA during pregnancy and breastfeeding. Searches were performed using the PubMed database from the National Library of Medicine—National Institutes of Health. The following keywords were used for the literature search: maternal obesity and PUFA metabolism and fetal development; maternal obesity and PUFA metabolism and breastfeeding; maternal obesity and PUFA metabolism and placenta.

3. Maternal Obesity and Its Consequences for the Offspring

Maternal pre-pregnancy obesity is independently associated with fetal overgrowth and with total body adiposity, abdominal fat accumulation, and lower fat free-mass in neonates [23,24,25,26,27,28]. These features are aggravated by an excessive GWG during mid and late pregnancy at the time of higher fat accretion in the fetus [25,29,30]. Moreover, high adiposity in neonates is related to insulin resistance, hyperinsulinemia, and a pro-inflammatory status characterized by high circulating levels of leptin and interleukin-6 (IL-6) [31,32]. In addition to these features, the intrahepatic fat content of neonates during the first week after birth is correlated with maternal body mass index (BMI) in obese women with gestational diabetes [33,34].
The effects of maternal obesity on offspring are maintained during childhood and even adulthood, such that it has been reported that obese women at the highest range of GWG confer a high risk of obesity to their offspring at 5 years old, and this higher risk is maintained during adolescence [35]. Moreover, maternal overweight during early pregnancy increases the risk factors for metabolic and cardiovascular diseases in adolescents regardless of their current BMI [36]. Interestingly, in a large study based on a Swedish nationwide register, the authors showed that the hazard risk for cardiovascular disease in offspring aged between 1 and 25 years old born to obese pregnant women ranged from 1.10 (95% CI 0.97–1.25) for overweight to 2.51 (95% CI 1.60–3.92) for obesity grade III [37]. In this regard, mice born to high fat diet (HFD)-induced obese dams had hyperphagia, physical inactivity, and altered adipocyte metabolism, leading to high adiposity and obesity. These results suggest that the metabolic changes associated with maternal obesity are programmed during prenatal life and maintained throughout the lifespan [38].
It is important to note that excessive GWG during pregnancy is a determinant factor on the offspring’s risk of developing metabolic disorders. Observations from the Generation R study conducted in the city of Rotterdam in the Netherlands showed that a higher weight gain in early, but not in mid or late pregnancy, was associated with increased risks of childhood overweight and clustering of cardio-metabolic factors at 6 years old [39]. Moreover, the effects of maternal pre-pregnancy BMI on childhood BMI, waist circumference, subcutaneous adipose tissue, and HDL-cholesterol at 10 years of age were attenuated by a reduced weight gain during pregnancy, which was independent of the current diet and physical activity of the child [40]. Regarding hepatic fat accumulation, adolescents born to mothers with a BMI over 30 kg/m2 before pregnancy had an increased hepatic fat fraction independent of gestational diabetes mellitus (GDM) and children adiposity [41]. Together with these metabolic and cardiovascular risk factors, it has been reported that children and adolescents born to mothers with obesity and/or excessive GWG present poor neurodevelopmental outcomes, with greater risk of attention deficit-hyperactivity disorder, autism spectrum disorder (ASD), developmental delay, and emotional-behavioral problems [42]. Furthermore, excessive weight and obesity during pregnancy have been associated with increased rates of malformations of the nervous system and epilepsy, which were worsened by the presence of neonatal hypoglycemia, respiratory distress syndrome, and neonatal jaundice [43].
Interestingly, the risk of developing overweight in children from 2 to 14 years of age born to women with maternal pre-pregnancy obesity was aggravated by the lack of breast-feeding with an OR of 6.1 (95% CI 2.9–13.1) when the two factors were present [44]. Similarly, a mice model of gestational and lactation overnutrition showed that the offspring had a severe metabolic phenotype in adulthood when they were exposed to HFD, and this was characterized by an increase in weight gain and adiposity, in addition to glucose homeostasis disturbances and lipid accumulation in the liver [45].
Some studies indicate that the offspring risk of developing metabolic, cardiovascular, and mental disorders differs between males and females. In this regard, at the neonatal period only, girls born to mothers with obesity exhibited increased skinfold thickness and serum leptin concentrations [46]. On the other hand, another study showed that at ages between 2 and 6 years old, boys born to mothers with obesity have higher body fat compared to those born to normal and overweight mothers, whereas no differences were observed in girls [47]. However, in adult women, the risk of developing type 2 diabetes increased according to the BMI of their mother during pregnancy [48]. The development of non-alcoholic fatty liver disease (NAFLD) was also associated with maternal obesity and higher GWG at early-mid pregnancy only in female offspring [49]. On the other hand, boys whose mothers were severely obese have shown more behavioral problems than those born to normal-weight mothers [50]. One study showed that adequate breastfeeding protects against adiposity excess only in boys born to overweight mothers [51], while another study showed that the clustering of cardiometabolic risk factors associated with maternal obesity is not influenced by breastfeeding [52]. In summary, maternal obesity during pregnancy and breastfeeding is associated with metabolic, cardiovascular, and behavioral effects in the offspring which might be sex-dependent, although the effects and consequences are highly heterogenous and not completely clear between males and females. Further studies are needed to clarify these issues.

4. Effects of Obesity on FA Metabolism during Pregnancy

4.1. Fatty Acid during Pregnancy

Maternal body fat accumulation during early pregnancy allows for the storage of important amounts of FA, especially LCPUFA derived from diet and maternal metabolism [53]. During the last trimester of gestation, the accumulation of fat depots in maternal tissues declines because of higher lipolysis and mobilization of triacylglycerols (TAG). The elevations of serum TAG, very low-density lipoprotein (VLDL), and non-esterified fatty acids (NEFA) increase FA bioavailability to be transferred to the fetus [53,54,55,56]. These metabolic changes are a consequence of pregnancy and associated with modifications in the synthesis and secretion of human placental lactogen, placenta growth hormone, cortisol, progesterone, estrogens, and adipokines derived from adipose tissue [57]. In this regard, in pregnant rats, estradiol levels are strongly correlated with the mRNA expression of Fads2 mainly during the mid-pregnancy, whereas serum progesterone concentrations are associated with the hepatic content of LCPUFA intermediates, particularly n-6 docosapentaenoic acid (C22:5n-6, DPA n-6) [58]. Taken together, these results suggest a placental control in maternal LCPUFA metabolism. In the maternal bloodstream, FA are contained in different proportions in TAG, phospholipids, and cholesterol esters or circulate as NEFA [56]. The FA composition in maternal erythrocytes indicates that the metabolism of SFA, MUFA, and PUFA is modified throughout pregnancy, showing an increase in palmitoleic acid (C16:1n-7, POA), nervonic acid (C24:1n-9, NA), n-3 LCPUFA such as n-3 docosapentaenoic acid (C22:5n-3, DPA n-3) and DHA, and n-6 LCPUFA such as LA and AA during pregnancy [59,60]. On the other hand, eicosapentaenoic acid (C20:5n-3, EPA) levels do not change significantly during gestation in erythrocytes or plasma. Interestingly, despite the increase in the absolute FA concentration during pregnancy, the ratio of essential PUFA in erythrocytes such as LA and ALA on total FA decreases, suggesting that pregnancy is associated with a reduction in their relative amounts in maternal circulation, probably due to the high transfer of these PUFA to the fetus [61]. In the serum, most of the palmitic acid (C16:0, PA) and oleic acid (C18:1n-9, OA) are found in TAG, whereas LA and DHA are found in phospholipids and TAG. This differential serum distribution between FA may influence their uptake and transfer [62]. Elegant in vivo studies that track FA metabolism with stable isotope FA tracers administered orally to pregnant women 4 to 12-h before labor have demonstrated a preferential transfer of DHA over other FA, such as PA, OA, and ALA [62,63]. In general, this selective enrichment of LCPUFA, mainly DHA and AA, in the fetal circulation during pregnancy and lactation is known as biomagnification [61,64]. These findings have given rise to the hypothesis that DHA and AA have to be transferred preferentially across the placenta to support their rapid accretion in the fetal nervous tissue during the period of brain growth spurt [65].
The dietary FA recommendations during pregnancy are shown in Table 1. According to the committee of the United Nations Food and Agriculture Organization and the World Health Organization (WHO), PUFA intake during pregnancy and lactation should represent between 20 and 35% of the total fat intake [66]. The main emphasis must be on dietary DHA and AA requirements because the risk of deficiency of these PUFA increases during pregnancy due to fetal neurodevelopment requirements [67]. Therefore, several organizations including the Food and Agriculture Organization, the WHO, the Perinatal Lipid Nutrition Project (PeriLip), and the Early Nutrition Project (EARNEST) of the European Commission recommend a daily intake of at least 300 mg of EPA and 200 mg of DHA per day or the consumption of one to two portions of fatty fish per week both in pregnant and lactating women [66,68]. Moreover, it has been indicated that the daily intake of AA should be around 800 mg [66]. The Food and Nutrition Board of the Institute of Medicine recommends a daily intake of LA and ALA during pregnancy at around 13 g per day and 1.4 g per day, respectively [69], although there is no clear consensus on the dietary requirements of these FA during pregnancy.
The beneficial effects of DHA during pregnancy or early postnatal period have been widely documented, showing a reduction in the risk of preeclampsia, preterm birth, and depression in the mothers, in addition to increases in birthweight and improvements in brain and visual development [72,73]. In support of these findings, a study by Much et al. on healthy pregnant women that received a fish oil supplement with 1200 mg n-3 LCPUFA (1020 mg DHA and 180 mg EPA), starting at week 14 of gestation and until the end of lactation, showed that maternal erythrocyte concentrations of n-3 LCPUFA at week 32 of gestation were positively associated with the children weight and length at birth [74]. On the other hand, AA and n-6 LCPUFA have been negatively correlated with the BMI and ponderal index at one year of life [74]. A randomized double-blind clinical trial conducted in 144 pregnant women found that infants of mothers who received 200 mg of DHA during pregnancy had a lower BMI at 21 months of age but no difference at 6 years old, suggesting that the effects of maternal DHA disappear as children grow older [75,76].
Pregnant women with overweight or obesity have an altered lipid profile characterized by elevated TAG, total cholesterol, LDL-cholesterol, and VLDL cholesterol, but lower HDL-cholesterol in comparison to women with normal weight [77,78,79,80]. In a small sample of participants, Scifres et al. observed that maternal serum levels of AA, both at the first and late second trimester of gestation, are higher in overweight/obese women in comparison to normal-weight women [60]. On the other hand, a large cohort study involving 5636 women at mid pregnancy described high SFA and low n-3 PUFA serum levels in women with pre-pregnancy obesity. Excessive GWG increases the serum levels of SFA, MUFA, and n-6 PUFA such as dihomo-γ-linolenic acid (C20:3n-6, DGLA) and AA, but decreases LA [81]. However, the FA profile in red blood cells showed lower levels of MUFA and LA but higher levels of DHA in women with maternal obesity [82]. Interestingly, a metabolomic analysis showed that the serum levels of NEFA n-6 PUFA and phosphatidylcholines containing DGLA were positively correlated with pre-pregnancy BMI [83].
Although changes in the blood FA profile seem to be determined mainly by maternal diet, obesity during gestation can also provoke modifications in PUFA metabolism in the maternal liver, affecting PUFA circulating levels (Figure 2). For instance, women with gestational obesity have a high prevalence of NAFLD, which might be associated with a reduction in liver PUFA synthesis due to a reduction in the activity of FADS1 and FADS2 as a result of an abnormal fat accumulation and oxidative stress [84,85,86,87]. The latter is a product of an elevated production of reactive oxidant molecules that exceeds the capacity of the cell’s antioxidant defense mechanisms [88]. Pregnancy is characterized by a state of oxidative stress that increases progressively from the late first trimester onward [89] and gestational obesity seems to aggravate this condition. For example, in obese rats there is an increase of reactive oxidative species (ROS) in the maternal liver compared to normal weight rats [90]. Importantly, according to the antioxidant and anti-inflammatory properties of n-3 PUFA in adults, it has been suggested that the maternal consumption of these FA could limit oxidative damage during pregnancy [91].

4.2. Placental Fatty Acid Transport

The placenta is responsible for the maternal-fetal transfer of oxygen, carbon dioxide, water, and all the nutrients necessary for the development of the fetus [92]. The exchange function occurs in placental chorionic villi that are composed of the syncytiotrophoblast, a multinucleated and polarized epithelium that comes from the fusion of mononuclear trophoblasts and is considered the transporter epithelium of the placenta [93]. The apical membrane has microvilli that are in contact with the maternal blood, and the basal membrane is in contact with the fetal capillaries. Because only very small substrates can cross both membranes of the syncytiotrophoblast, macro and micronutrients are transported through nutrient transporters that are highly expressed in this epithelium [94].
In the placenta, FA are used for its growth and development or mobilized to the fetus mainly during mid and late pregnancy. Maternal TAG, by the action of placental lipases, are hydrolyzed to NEFA. Then, the placenta can uptake and move the NEFA across the membrane through a flip-flop mechanism or with the help of binding proteins and transporters (Figure 3) [95]. These include the placenta plasma membrane’s fatty acid binding protein (p-FABPpm), fatty acid translocase (FAT/CD36), and fatty acid transport proteins (FATP) [65]. Functionally, p-FABPpm is exclusively expressed in the apical membrane and exhibits a high affinity for LCPUFA, suggesting that this protein is involved in the preferential uptake of these PUFA. FAT/CD36 is a class B scavenger receptor protein, located both at the apical and basal membranes, and it is involved in angiogenesis, atherosclerosis, inflammation, and lipid metabolism. FATP are integral membrane proteins important for the cellular uptake of LCPUFA. The FATP family consists of at least six related members, of which five are expressed in the human placenta (FATP1-4, and 6; SLC27A1-4, and 6). It is important to note that these FATP do not have specificity for a certain type of FA. Fatty acid binding proteins (FABP) are cytosolic proteins allowing for the trafficking of FA to sites within syncytiotrophoblast for esterification, β-oxidation, or transfer to the fetus. The human placenta expresses four different isoforms of FABP: FABP1, FABP3, FABP4, and FABP5 (FABP1, 3, 4, and 5) [94,96]. In addition, it was recently established that the placenta expresses the major facilitator superfamily domain-containing protein 2 (MFSD2a) [97], which is a sodium-dependent lysophospholipid transporter known for its expression at the blood brain barrier. MFSD2a uptakes DHA in the form of lysophosphatidylcholine (LPC) [98]. Interestingly, the placental expression of MFSD2a is correlated with DHA levels in cord blood, suggesting its role in the maternal-fetal transfer of DHA [97].
In pregnant women, excessive fat mass increases the circulating levels of pro-inflammatory cytokines, IL-6, and tumor necrosis factor- α (TNF-α), among others [99], which further accentuates pregnancy-induced insulin resistance. These pro-inflammatory molecules impair, mainly during the second and third trimester of gestation, multiple metabolic pathways that result in higher maternal circulating levels of TAG, LDL-, and VLDL-cholesterol [60]. Moreover, it has been observed that obesity promotes a lipotoxic placental environment with macrophage infiltration and high levels of IL-1, TNF-α, IL-6, and oxidative stress, in addition to alterations in energy metabolism [21]. Cytokines regulate the expression, activity, and subcellular localization of placental FA transporters. In this regard, it has been shown that IL-6 stimulates trophoblast FA accumulation [100]. The effect of oxidative stress on the placental transport of FA is unknown, but it has been hypothesized that it could produce an increase in pro-inflammatory cytokines, thus indirectly regulating placental lipid metabolism (Figure 1) [101]. On the other hand, few studies have pointed out the role of maternal PUFA consumption on placental inflammation and oxidative stress. Both in human and rodent placenta, supplementation with n-6 PUFA does not seem to modify the inflammation that occurs at the term of pregnancy, although it slightly elevates the gene expression of pro-inflammatory cytokines such as TNF-α, IL6, and IL1-β. On the other hand, n-3 PUFA supplementation increases resolvin and protectin levels, which are derived from DHA and EPA and have anti-inflammatory properties [102,103]. Moreover, n-3 PUFA intake during pregnancy reduces placental oxidative damage due to an increase in antioxidant capacity in the placental labyrinth zone at the final third of rat pregnancy [104]. Similar findings have been reported in placental BeWo cells and in human placental explants [105]. Interestingly, in vitro models have shown that high levels of DHA (100 µM) enhance lipid peroxidation and DNA oxidative damage in placental cells and explants [105]. On the other hand, pregnant rats fed with a LA-rich diet showed decreased placental concentrations of IL-7 and IL-10, indicating that this n-6 PUFA has both pro-inflammatory and anti-inflammatory properties in the placenta [106]. DHA seems to have beneficial effects on fetal growth but no effects on placental tissue in rat models. On the other hand, maternal AA levels were positively correlated with placental lipid peroxidation and fetal growth restriction, mainly in female fetuses, indicating a sex-dependent effect [107]. It is interesting to note that there is evidence suggesting that fetal PUFA demands vary in a sex-dependent way. In pregnant sows, García-Contreras showed that the male fetus has higher ratios of MUFA to SFA and C18:1 to C18:0, indicating higher stearoyl-CoA desaturase activity in the liver [108]. Moreover, male fetuses had a higher n-6/n-3 ratio than females [108].
Experimental evidence in ovine models indicates that maternal obesity alters placental FA transport through changes in transporter levels rather than TAG [109]. On the other hand, studies in human placenta have found similar findings but have been inconsistent regarding the expression of different transporters and binding proteins. Hirschmugl et al. observed that maternal obesity significantly modified the expression of placental genes related to the transport and storage of neutral lipids (ATGL, FATP1, FATP3, PLIN2, PPARG, and CGI-58), resulting in a higher content of placental TAG [110]. Dubé et al. found an increase in mRNA levels and protein expression of FAT/CD36, while both FABP1 and FABP3 were found to be decreased [111]. Lager et al. showed that FATP2 expression was correlated with maternal BMI in contrast to FAT/CD36 and FATP4 [112]. Interestingly, Calabuig-Navarro et al. observed a higher esterification and storage of lipids, lower mitochondrial fatty acid oxidation, but elevated peroxisomal fatty acid oxidation in human placenta from mother with obesity compared to from normal-weight mothers [113]. The authors concluded that these changes suggest an adaptative mechanism that limits the transfer of FA to the fetus in case of maternal obesity [113].
Evidence from studies that have measured FA in erythrocytes from cord blood in women with gestational obesity indicates that the modifications in FA transporters and placental uptake are reflected by an altered FA profile in the fetal circulation characterized by decreased PUFA, both total n-6 and DHA (Figure 2) [82]. A recent study that looked at the placental transport of 13C-PUFA in pregnant women with obesity described a higher concentration of LA and DHA in the maternal plasma NEFA fraction of mothers with obesity. Nevertheless, the umbilical cord venous to maternal plasma ratio of 13C-LA was lower, whereas 13C-DHA tended to be lower in women with obesity compared to controls, suggesting a reduced maternal-fetal transfer of these FA [21]. However, Draycott et al. observed that fetal 13C-LA concentrations seem to be influenced by the total amount of FA in the maternal diet, more so than by the n-6 to n-3 ratio, concluding that a higher amount of total FA consumed would favor the placental transport of 13C-LA [114]. In addition, Shrestha et al. [14] observed that the exposure to LA increased the expression of the transporters FATP1, FATP4, and FABP5 and decreased that of FABP3 in a culture of cytotrophoblasts, indicating that these proteins could regulate the transport of this PUFA. On the other hand, studies on the placental transport of 13C-LA are controversial, and it is still unclear whether maternal-fetal transfer is affected by obesity. Although a higher LA uptake has been reported in isolated cytotrophoblasts from women with obesity, this effect is not maintained after the differentiation into syncytiotrophoblast [111]. On the other hand, a preliminary study of a placental metabolomic analysis of obese women showed an increase in PA concentration and a decrease in DHA, AA, and stearic acid (C18:0, SA) [115]. However, these results could be influenced by the presence of GDM. Placenta explants from obese women did not show any differences in the uptake of AA and DHA, but those from women with obesity and male fetus showed an increased uptake of OA [116].
In general, changes in the placental transfer of FA induced by maternal obesity could lead to alterations in different fetal tissue, such as the liver, adipose tissue, muscle, and brain. In humans, pre-pregnancy obesity is associated with increased oxidative stress in the newborn; a study in pregnant women showed a direct relationship between umbilical cord malondialdehyde levels, a biomarker of oxidative stress, and maternal BMI [117]. With regards to n-3 PUFA, it has been observed that, depending on experimental conditions, the administration of n-3 PUFA can have both pro and antioxidant effects [91,118]. Consistently, Shoji et al. showed that modest levels of DHA alleviate oxidative DNA damage, whereas high levels of DHA accelerate lipid peroxidation [118].

5. Impact of Maternal Obesity on PUFA Metabolism and Fetal Development

Maternal obesity alters the structure and function of the placenta, leading to modifications in the PUFA delivered to the fetus, producing changes in fetal development that can result in metabolic disturbances within multiple organs (Figure 1) [119,120]. Once FA are transferred to the fetal circulation, they are mobilized to tissue binding through α-fetoprotein [121]. In baboons, visceral fetal organs such as the liver, kidney, and lung have the capacity to synthesize AA from LA. In the heart, on the other hand, LA is more important for oxidation than for AA synthesis [122]. One study that analyzed fetal liver microsomes showed FADS1 and FADS2 activities in human fetal liver as early as 18 and 22 weeks of gestation, but these are lower than in other species, especially rodents [123,124]. Therefore, fetal organs have the capacity to produce DHA from ALA and AA from LA, but these capacities appear to be limited, suggesting that DHA and AA must also be obtained from the mother [125].
Although the gene expression of the enzymes required for de novo lipogenesis and fatty acid oxidation (FAO) have been reported in the fetal liver, it is considered that in this organ FA catabolism is suppressed in comparison to what happens during the postnatal period, as the major source of energy supply for the fetus comes from carbohydrates and amino acids rather than fat [126]. However, there is evidence showing a high rate of lipogenesis in slices from murine fetal liver compared to those from adult liver [5,127]. Measurements in murine and human tissues indicate that the activity of enzymes involved in FAO and the presence of different acylcarnitines have also been reported, indicating that these processes are functional in the fetal liver [128,129].
Studies performed in primates and mice have shown that HFD-induced maternal obesity has been associated with lipid accumulation in the fetal liver, suggesting that an abnormal supply of FA could alter the fetal hepatic FA metabolism, and this could be an early mechanism contributing to the origin of NAFLD [130,131]. It is interesting to note that non-human primate models do not develop obesity during pregnancy but show fat accumulation in the fetal liver, suggesting a dietary role in this process. Gene expression of de novo lipogenesis enzymes, such as fatty acid synthase (FAS) and sterol regulatory element-binding protein 1 (SREBP-1), is increased in the fetal liver of non-human primates fed with an obesogenic diet [126]. On the other hand, rats with gestational obesity induced by a diet rich in saturated fat exhibited a decreased gene expression of FAO enzymes [132]. Together with these changes, alterations in the plasma FA profile have been found showing reduced levels of n-6 DPA concomitantly with elevated plasma levels of LA and AA (Figure 2) [133]. Interestingly, the consumption of a HFD enriched with dietary n-3 PUFA, such as ALA, EPA and DHA in dams leads to a reduction in TAG accumulation in the fetal liver, suggesting that the maternal diet composition is fundamental in the prevention or worsening of this condition [134].
Maternal supplementation with n-3 PUFA during pregnancy decreases SREBP-1 protein expression and increases peroxisome proliferator-activated receptor alpha (PPAR-α), which promotes FAO in the offspring’s liver at 3 days of age [135]. Maternal obesity can also produce a mitochondrial dysfunction, evidenced by impairments in the fetal hepatic respiratory chain capacity, which could contribute to increased oxidative stress [132,136,137,138]. However, there is no evidence about how PUFA can regulate the ROS levels in fetal liver, which could be of interest in order to avoid early damage in the liver.
PUFA are stored in fetal adipose tissue, such that at the end of pregnancy the amounts of DHA and AA are several times higher in fetal than in maternal adipose tissue [139]. Moreover, LCPUFA act on preadipocytes, promoting adipogenesis due to transcriptional expression regulation of genes related to lipid metabolism, such as those encoding for the PPAR family [5]. In maternal obesity, evidence from animal and human studies supports the hypothesis that an imbalance of essential PUFA at early stages of pregnancy alters fetal adipose tissue development, favoring an abnormal lipid metabolism [140]. Similarly, the excess and imbalance of PUFA in fetal circulation induce alterations of skeletal muscle development, resulting in changes in its function and metabolism that predispose the child to future metabolic disorders. A study conducted in sheep suggests that maternal obesity induced by HFD generates modifications in fetal myogenesis and in various signaling cascades, such as the insulin signaling pathway, adipogenesis, and oxidative stress pathways [141]. Another group found a reduced muscle fiber density that could be associated with an increase in adipogenesis (diversion from myogenesis to adipogenesis) and lipid accumulation in fetuses of obese pregnant sheep [142]. Interestingly, studies in cell lines have shown that EPA and DHA reduce myogenesis and increase adipogenesis in myotube formation [143].
PUFA are central in fetal brain development. In this regard, DHA is particularly important because it regulates cell survival, neuroinflammation, and neurogenesis and participates in signal transduction and blood-brain barrier (BBB) permeability [13,144]. AA is crucial for several functions in the brain, such as neuronal firing, signaling, and long-term potentiation. Moreover, AA contributes to the maintenance of hippocampal plasticity in part because it activates the PPARγ [145].
The absorption of esterified lipids through the BBB is essential for the accumulation of DHA, in which its dissociation from albumin lysophospholipids or the release of circulating lipoproteins by lipases of brain endothelium are required to cross to the luminal plasma membrane of the BBB by passive diffusion or facilitated transport [144]. FA bind FABP for trafficking within the cell or are trapped by the activation to acyl-CoA catalyzed by the activity of FATP [13,144]. FABP7 levels correlate with neuronal differentiation, whereas FABP5 influences endothelial cells within the brain. Recently, it has been proposed that brain uptake of DHA occurs mostly through MFSD2a, which uptakes lysophospholipids at the BBB [144,146], although this is still controversial.
Hypothalamic inflammation induced by HFD has been shown to deregulate energy homeostasis, leading to insulin resistance, glucose intolerance, and obesity [147,148]. There are several possible mechanisms that can induce hypothalamic inflammation through the activation of different intracellular processes in the hypothalamic glial cells, including oxidative stress, endoplasmic reticulum stress, RNA stress, autophagy defects, or the activation of toll-type receptors (TLRs) and cytokines. Most of these intracellular processes converge in the activation of the N-terminal kinase c-Jun (JNK) and the IκB kinase-nuclear factor kappa B (IKK/NF-κB) [148,149,150,151]. It has been observed in rats that maternal obesity induces an upregulation of several members of the toll-like receptor 4 signaling cascade and the subsequent activation of inflammatory pathways in the hypothalamus [152].

6. Impact of Maternal Obesity on PUFA Metabolism during Breastfeeding and Neonatal Development

Clinical trials have shown that the composition of breast milk appears to be determined through adaptations that occur during pregnancy with the aim of supporting the newborn according to the gestational age [153]. During pregnancy, prolactin in concert with glucocorticoids and insulin are the main inductors of breast development and differentiation. Twenty-four hours after birth, the drop of progesterone along with glucocorticoids induce cellular and molecular changes in the mammary epithelium, resulting in modifications of key pathways of carbohydrate and lipid metabolism, in addition to increasing protein synthesis and milk production [154]. In turn, prolactin induces the expression of lipogenic genes involved in de novo FA lipogenesis and β-oxidation in the mammary gland both in humans and mice [155,156].
Breast milk composition can be influenced by maternal age, weight, diet, and health condition [157]. Moreover, it varies throughout the lactation period. First, milk corresponds to colostrum produced until the 5th day of lactation, then transition milk, until the 14th day, and finally mature milk [157]. Colostrum is characterized by a high content of protein (23 g/L) and immunological factors. This milk is rich in n-6 and n-3 PUFA, specifically LA and ALA, and is the initial source of LCPUFA to breastfed newborns [158]. LCPUFA composition of mature human milk in lactating women from Europe, Africa, Asia, and South America shows that the major LCPUFA in human milk are AA, DGLA, and eicosadienoic acid (C20:2n-6, EA) from n-6 series in addition to DHA and DPA from n-3 series [159]. Overall, the concentration of FA in breast milk increases from the first to the fourth month of lactation and then remains stable until the 6th month. However, n-3 LCPUFA and DHA decrease from the 4th month of lactation, resulting in a high n-6 to n-3 LCPUFA ratio [19]. Although DHA content in breast milk varies in relation to the mother’s DHA intake, a global average of DHA in breast milk has been estimated at 0.32 ± 0.22% [160]. Interestingly, DHA supplementation was shown to slightly reduce n-6 LCPUFA levels in breast milk, which reduces the n-6: n-3 ratio, suggesting that DHA could modify the transport or metabolism of n-6 LCPUFA [161].
Maternal obesity impacts breastfeeding because the excess of adipose mass affects hormonal regulation and alters the composition of breast milk (Figure 1). In this regard, obese women had reduced levels of prolactin compared to women with normal weight [162]. In addition, insulin resistance and impaired insulin levels, which are closely related to obesity, delay the onset of lactogenesis and alter milk production [163]. These effects are mediated, in part, by an abnormal de novo lipogenesis related to an increased AMP-activated protein kinase (AMPK) activity, the inhibition of acetyl-CoA carboxylase, and a decreased PUFA synthesis in the mammary gland [164]. This evidence explains why women with obesity breastfeed for a shorter duration and introduce complementary food earlier than women with normal weight [165]. Moreover, a study conducted by Lima et al. in pregnant rats divided into two groups, one control and one fed with a HFD, showed that maternal lipids are reflected in milk composition, and a HFD leads to hypercholesterolemia and visceral fat accumulation in the offspring [166]. Maternal obesity is also associated with the alteration of immunological factor concentrations in human milk, such as C-reactive protein (CRP), leptin, IL-6, insulin, TNF-α, ghrelin, adiponectin, and obestatin. Moreover, excessive weight and obesity appear to have a clear influence on the immunological properties of human milk [167]. The alterations of bioactive properties in the human milk of obese mothers could increase the incidence of obesity, insulin resistance, type 2 diabetes, and other adverse metabolic outcomes in the offspring [168]. Despite this, there is not enough evidence to suggest that these changes should preclude infant’s breastfeeding, because the benefits of maternal milk are well established [169].
In the colostrum of women with obesity, higher levels of LA and AA and lower levels of ALA and DHA have been observed compared to normal-weight mothers [10]. The FA profile of mature breast milk from mothers with obesity is characterized by: (i) increased SFA levels, mainly PA; (ii) decreased MUFA, such as OA; (iii) lower n-3 PUFA, including ALA, DPA, EPA, and DHA; (iv) higher n-6 PUFA, DGLA, and adrenic acid (C22:4n-6, ADA); and (v) elevated n-6 to n-3 ratio [168,169,170,171,172,173,174,175]. Similarly, the breast milk of Japanese monkeys fed with a HFD exhibited lower levels of EPA and DHA, along with a higher n-6 to n-3 ratio than those fed with a control diet [176]. Interestingly, there are other types of lipids with anti-inflammatory and anti-diabetics properties in human breast milk, like palmitic acid and hydroxystearic acids, which were found to be lower in the milk of obese mothers compared to normal-weight mothers [177]. In general, these modifications can be attributed to differences in the mobilization of endogenous FA stores, synthesis in the maternal liver and breast tissue, or dietary differences [174]. In this sense, a maternal diet with low levels of essential PUFA results in a decrease of PUFA levels in breast milk (Figure 2) [168].
The content of PUFA from n-3 and n-6 series in breast milk has been associated with modifications in weight and fat distribution in the infant. It has been established that the colostrum content of AA and DHA is inversely correlated with infant BMI at 6 months of age, whereas the n-6 to n-3 ratio is positively associated with BMI. Moreover, in infants of overweight mothers, DHA and n-3 LCPUFA levels in colostrum were positively associated with cognitive scores, while the n-6 to n-3 ratio was inversely associated with it. Moreover, SFA, as well as the ratio of unsaturated to SFA in mature breast milk at 3 months, are correlated with children’s weight gain at 13 months [169]. In turn, a positive association between ALA levels in mature milk with cognition has been observed in infants born to mothers with obesity [170].
Several studies have shown that the type of dietary FA consumed by dams during pregnancy and/or lactation can have beneficial or adverse consequences on the health of their offspring (Figure 2). The maternal consumption of diets rich in trans FA and SFA can lead to impairments in the metabolism and development of the offspring, basically by TLR4 activation and disturbances in glucose and lipid homeostasis. However, the majority of the studies evaluating the early exposition to PUFA, in particular n-3 PUFA, have shown benefits in the offspring development and epigenetic regulation, which seem to play a role in the prevention of obesity, insulin resistance, and the risk of developing cardiovascular diseases later in life. In relation to MUFA, limited evidence indicates that maternal intake can stimulate thermogenic capacity and change liver metabolism, favoring the offspring’s health [178]. In animal models and clinical trials based on n-3 PUFA supplementation during pregnancy and lactation, researchers have demonstrated the importance of these FA during breastfeeding. However, the results of these studies are heterogenous. A study in mice with maternal EPA + DHA-rich fish oil supplementation potentiated the development of fetal brown adipose tissue, a crucial regulator of energy expenditure that reduces the susceptibility to develop obesity [179,180]. On the other hand, male and female rat offspring born to dams supplemented with a diet enriched with EPA and DHA exhibited large subcutaneous fat depots without effect on the expression of adipogenic or lipogenic genes in adipose tissue at early postnatal life [178,181].
In women with maternal obesity, Rouille et al. showed a positive relationship between the maternal ratio of n-3 PUFA to total PUFA intake and fetal growth in overweight pregnant women [182]. Likewise, a randomized controlled trial involving 556 pregnant women observed that women with obesity supplemented from mid-pregnancy until delivery with 2 g/d of marine n-3 PUFA containing 800 mg of DHA plus 1200 mg of EPA had an attenuated increment in plasma n-3 concentrations and a lower reduction of n-6 to n-3 ratio compared to women with normal weight [183]. The serum concentration of n-3 PUFA was positively associated with the sum of skinfold at one year of age in women’s offspring supplemented with DHA during pregnancy and lactation [74]. Likewise, DHA and n-3 PUFA of early breast milk were positively related to the proportion of subcutaneous and preperitoneal fat (subcutaneous/preperitoneal) assessed by ultrasound in offspring at 6 weeks postpartum [184]. Similar observations were made in a Spanish cohort in which the authors found that n-3 PUFA were positively related to weight and the percentage of lean mass in the infants [5], but no associations were found in infants of mothers with overweight or obesity [5].
There is evidence indicating that supplementation with n-3 PUFA can improve some neurodevelopment parameters. In a study in which the school menu was modified to offer fish twice a week to 8–10 year-old healthy children, the authors report a significantly improved school performance and reading comprehension, in addition to increases in EPA and DHA levels [185].
Similarly, Richardson et al. conducted a randomized controlled trial in 5–12 year-old children with developmental coordination disorders and observed that daily supplementation with n-3 PUFA (558 mg of EPA and 174 mg of DHA), n-6 PUFA (60 mg of γ-linoleic acid), and vitamin E (9.6 mg of α-tocopherol) resulted in improvements in reading, spelling, and behavioral scores [186]. These data suggest that a supplementation with n-3 PUFA could improve some of the parameters that cause neurodevelopmental abnormalities associated with gestational obesity.

7. Conclusions

Maternal obesity and excessive GWG are strongly associated with changes in the FA profile of maternal and fetal circulation during pregnancy and with breast milk composition during lactation. Moreover, the pro-inflammatory status induced by obesity alters placental FA transporters, resulting in an imbalance in the transport of nutrients, including FA. Additionally, maternal diets and hormonal and metabolic modifications affect breastfeeding and breast milk composition. In general, obesity leads to modifications in PUFA levels in breast milk and fetal circulation with lower n-3 PUFA (mainly EPA and DHA) and higher n-6 PUFA (mainly LA and AA) resulting in a high n-6 to n-3 ratio. Early nutrition patterns in fetuses and neonates can influence adiposity, in addition to the risk of developing metabolic and behavioral disorders, and these associations can persist throughout the children’s whole life.

Author Contributions

Conceptualization, M.M. and R.V.; writing original draft, D.Á., Y.M., M.O., M.M. and R.V.; writing review and editing, M.M., R.C.-W. and R.V. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Acknowledgments

This work was supported by Fondo Nacional de Desarrollo Científico y Tecnológico (National Fund for Scientific and Technological Research), FONDECYT Grant 1181798 (MM).

Conflicts of Interest

The authors declare no conflict of interest.

Abbreviations

AAarachidonic acid
ADAadrenic acid
AL α–linolenic acid
AMP AMP-activated protein kinase
ASDautism spectrum disorder
ATG adipose triglyceride lipase
BBBblood-brain barrier
BMIbody mass index
CGI-58comparative gene identification-58
CRP C-reactive protein
DGLAdihomo-γ-linolenic acid
DH docosahexaenoic acid
DPAdocosapentaenoic acid
EARNESTEarly Nutrition Project
ELOVLelongation of very long chain fatty acids protein
EA
EPA
eicosadienoic acid
eicosapentaenoic acid
FABP
FADS
fatty acid binding protein
fatty acid desaturase
FAO fatty acid oxidation
FASfatty acid synthase
FAfatty acid
FAT/CD36fatty acid translocase
FATPfatty acid transport protein
GDMgestational diabetes mellitus
GWG: gestational weight gain
HFD
IKK/NF-κB
high fat diet
IκB kinase-nuclear factor kappa B
IκBkinase-nuclear factor kappa B
ILinterleukin
JNkinase c-Jun
LAlinoleic acid
LCPUFAlong chain polyunsaturated fatty acid
MFSD2a main facilitator superfamily domain containing 2a
MUFAmonounsaturated fatty acid
NAFLDnon-alcoholic fatty liver disease
NA nervonic acid
NEFA non-esterified fatty acid
OAoleic acid
PApalmitic acid
PeriLipPerinatal Lipid Nutrition Project
p-FABPpmplacenta plasma membrane fatty acid binding protein
PLIN2perilipin 2
POApalmitoleic acid
PPAR-γperoxisome proliferator-activated receptor gamma
PPAR-αperoxisome proliferator-activated receptor alpha
ROS
SA
SFA
reactive oxidative species
stearic acid
saturated fatty acid
SREBP-1
TAG
sterol regulatory element-binding protein 1
triacylglycerols
TLRstoll-type receptors
TNF-αtumor necrosis factor-alpha
VLDLvery low-density lipoprotein

References

  1. Downs, D.S. Obesity in Special Populations: Pregnancy. Prim Care 2016, 43, 109–120. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  2. Mission, J.F.; Marshall, N.E.; Caughey, A.B. Pregnancy risks associated with obesity. Obstet. Gynecol. Clin. N. Am. 2015, 42, 335–353. [Google Scholar] [CrossRef] [PubMed]
  3. Valsamakis, G.; Kyriazi, E.L.; Mouslech, Z.; Siristatidis, C.; Mastorakos, G. Effect of maternal obesity on pregnancy outcomes and long-term metabolic consequences. Hormones (Athens) 2015, 14, 345–357. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  4. Kitsiou-Tzeli, S.; Tzetis, M. Maternal epigenetics and fetal and neonatal growth. Curr. Opin. Endocrinol. Diabetes Obes. 2017, 24, 43–46. [Google Scholar] [CrossRef] [PubMed]
  5. Herrera, E.; Desoye, G. Maternal and fetal lipid metabolism under normal and gestational diabetic conditions. Horm. Mol. Biol. Clin. Investig. 2016, 26, 109–127. [Google Scholar] [CrossRef]
  6. Feng, Y.; Jiang, C.-D.; Chang, A.-M.; Shi, Y.; Gao, J.; Zhu, L.; Zhang, Z. Interactions among insulin resistance, inflammation factors, obesity-related gene polymorphisms, environmental risk factors, and diet in the development of gestational diabetes mellitus. J. Matern. Fetal Neonatal Med. 2019, 32, 339–347. [Google Scholar] [CrossRef]
  7. Hernandez-Rodas, M.C.; Valenzuela, R.; Videla, L.A. Relevant Aspects of Nutritional and Dietary Interventions in Non-Alcoholic Fatty Liver Disease. Int. J. Mol. Sci. 2015, 16, 25168–25198. [Google Scholar] [CrossRef] [Green Version]
  8. Koletzko, B.; Godfrey, K.M.; Poston, L.; Szajewska, H.; van Goudoever, J.B.; de Waard, M.; Brands, B.; Grivell, R.M.; Deussen, A.R.; Dodd, J.M.; et al. Nutrition During Pregnancy, Lactation and Early Childhood and its Implications for Maternal and Long-Term Child Health: The Early Nutrition Project Recommendations. Ann. Nutr. Metab. 2019, 74, 93–106. [Google Scholar] [CrossRef]
  9. Desforges, M.; Sibley, C.P. Placental nutrient supply and fetal growth. Int. J. Dev. Biol. 2010, 54, 377–390. [Google Scholar] [CrossRef]
  10. Armand, M.; Bernard, J.Y.; Forhan, A.; Heude, B.; Charles, M.A.; Annesi-Maesano, I.; Botton, J.; Dargent-Molina, P.; de Lauzon-Guillain, B.; Ducimetière, P.; et al. Maternal nutritional determinants of colostrum fatty acids in the EDEN mother-child cohort. Clin. Nutr. 2018, 37, 2127–2136. [Google Scholar] [CrossRef]
  11. Gura, T. Nature’s first functional food. Science 2014, 345, 747–749. [Google Scholar] [CrossRef] [PubMed]
  12. Christian, P.; Mullany, L.C.; Hurley, K.M.; Katz, J.; Black, R.E. Nutrition and maternal, neonatal, and child health. Semin. Perinatol. 2015, 39, 361–372. [Google Scholar] [CrossRef] [PubMed]
  13. Bazinet, R.P.; Laye, S. Polyunsaturated fatty acids and their metabolites in brain function and disease. Nat. Rev. Neurosci. 2014, 15, 771–785. [Google Scholar] [CrossRef] [PubMed]
  14. Shrestha, N.; Cuffe, J.S.M.; Holland, O.J.; Perkins, A.V.; McAinch, A.J.; Hryciw, D.H. Linoleic Acid Increases Prostaglandin E2 Release and Reduces Mitochondrial Respiration and Cell Viability in Human Trophoblast-Like Cells. Cell. Physiol. Biochem. 2019, 52, 94–108. [Google Scholar] [CrossRef] [PubMed]
  15. Sattar, N.; Berry, C.; Greer, I.A. Essential fatty acids in relation to pregnancy complications and fetal development. BJOG Int. J. Obstet. Gynaecol. 1998, 105, 1248–1255. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  16. Guillou, H.; Martin, P.G.; Pineau, T. Transcriptional regulation of hepatic fatty acid metabolism. In Lipids in Health and Disease; Springer: Dordrecht, The Netherlands, 2008; pp. 3–47. [Google Scholar] [CrossRef]
  17. Renault, K.M.; Carlsen, E.M.; Nørgaard, K.; Nilas, L.; Pryds, O.; Secher, N.J.; Olsen, S.F.; Halldorsson, T.I. Intake of Sweets, Snacks and Soft Drinks Predicts Weight Gain in Obese Pregnant Women: Detailed Analysis of the Results of a Randomised Controlled Trial. PLoS ONE 2015, 10, e0133041. [Google Scholar] [CrossRef]
  18. Ainscough, K.M.; O’Brien, E.C.; Lindsay, K.L.; Kennelly, M.A.; O’Sullivan, E.J.; O’Brien, O.A.; McCarthy, M.; De Vito, G.; McAuliffe, F.M. Nutrition, Behavior Change and Physical Activity Outcomes From the PEARS RCT-An mHealth-Supported, Lifestyle Intervention Among Pregnant Women With Overweight and Obesity. Front. Endocrinol. 2019, 10, 938. [Google Scholar] [CrossRef]
  19. Barrera, C.; Valenzuela, R.; Chamorro, R.; Bascunan, K.; Sandoval, J.; Sabag, N.; Valenzuela, F.; Valencia, M.P.; Puigrredon, C.; Valenzuela, A. The Impact of Maternal Diet during Pregnancy and Lactation on the Fatty Acid Composition of Erythrocytes and Breast Milk of Chilean Women. Nutrients 2018, 10, 839. [Google Scholar] [CrossRef] [Green Version]
  20. Cetin, I.; Alvino, G.; Cardellicchio, M. Long chain fatty acids and dietary fats in fetal nutrition. J. Physiol. 2009, 587, 3441–3451. [Google Scholar] [CrossRef]
  21. Gazquez, A.; Prieto-Sanchez, M.T.; Blanco-Carnero, J.E.; Ruiz-Palacios, M.; Nieto, A.; van Harskamp, D.; Oosterink, J.E.; Schierbeek, H.; van Goudoever, J.B.; Demmelmair, H.; et al. Altered materno-fetal transfer of 13C-polyunsaturated fatty acids in obese pregnant women. Clin. Nutr. 2019, 39, 1101–1107. [Google Scholar] [CrossRef]
  22. Ballard, O.; Morrow, A.L. Human Milk Composition: Nutrients and Bioactive Factors. Pediatr. Clin. 2013, 60, 49–74. [Google Scholar] [CrossRef] [Green Version]
  23. Gaudet, L.; Ferraro, Z.M.; Wen, S.W.; Walker, M. Maternal Obesity and Occurrence of Fetal Macrosomia: A Systematic Review and Meta-Analysis. Biomed. Res. Int. 2014, 2014, 640291. [Google Scholar] [CrossRef] [PubMed]
  24. Ehrenberg, H.M.; Mercer, B.M.; Catalano, P.M. The influence of obesity and diabetes on the prevalence of macrosomia. Am. J. Obstet. Gynecol. 2004, 191, 964–968. [Google Scholar] [CrossRef] [PubMed]
  25. Ratnasiri, A.W.G.; Lee, H.C.; Lakshminrusimha, S.; Parry, S.S.; Arief, V.N.; DeLacy, I.H.; Yang, J.-S.; DiLibero, R.J.; Logan, J.; Basford, K.E. Trends in maternal prepregnancy body mass index (BMI) and its association with birth and maternal outcomes in California, 2007–2016: A retrospective cohort study. PLoS ONE 2019, 14, e0222458. [Google Scholar] [CrossRef] [Green Version]
  26. Whitelaw, A.G. Influence of maternal obesity on subcutaneous fat in the newborn. Br. Med. J. 1976, 1, 985–986. [Google Scholar] [CrossRef] [Green Version]
  27. Hull, H.R.; Dinger, M.K.; Knehans, A.W.; Thompson, D.M.; Fields, D.A. Impact of maternal body mass index on neonate birthweight and body composition. Am. J. Obstet. Gynecol. 2008, 198, 416.e1–416.e6. [Google Scholar] [CrossRef]
  28. Carlsen, E.; Renault, K.; Nørgaard, K.; Nilas, L.; Jensen, J.; Hyldstrup, L.; Michaelsen, K.; Cortes, D.; Pryds, O. Newborn regional body composition is influenced by maternal obesity, gestational weight gain and the birthweight standard score. Acta Paediatr. 2014, 103, 939–945. [Google Scholar] [CrossRef]
  29. Widen, E.M.; Factor-Litvak, P.R.; Gallagher, D.; Paxton, A.; Pierson, R.N., Jr.; Heymsfield, S.B.; Lederman, S.A. The Pattern of Gestational Weight Gain is Associated with Changes in Maternal Body Composition and Neonatal Size. Matern. Child Health J. 2015, 19, 2286–2294. [Google Scholar] [CrossRef] [Green Version]
  30. Ruchat, S.-M.; Allard, C.; Doyon, M.; Lacroix, M.; Guillemette, L.; Patenaude, J.; Battista, M.-C.; Ardilouze, J.-L.; Perron, P.; Bouchard, L.; et al. Timing of Excessive Weight Gain During Pregnancy Modulates Newborn Anthropometry. J. Obstet. Gynaecol. Can. 2016, 38, 108–117. [Google Scholar] [CrossRef]
  31. Lindsay, K.L.; Brennan, L.; Rath, A.; Maguire, O.C.; Smith, T.; McAuliffe, F.M. Gestational weight gain in obese pregnancy: Impact on maternal and foetal metabolic parameters and birthweight. J. Obstet. Gynaecol. 2018, 38, 60–65. [Google Scholar] [CrossRef]
  32. Catalano, P.M.; Presley, L.; Minium, J.; Hauguel-de Mouzon, S. Fetuses of Obese Mothers Develop Insulin Resistance in Utero. Diabetes Care 2009, 32, 1076–1080. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  33. Modi, N.; Murgasova, D.; Ruager-Martin, R.; Thomas, E.L.; Hyde, M.J.; Gale, C.; Santhakumaran, S.; Doré, C.J.; Alavi, A.; Bell, J.D. The Influence of Maternal Body Mass Index on Infant Adiposity and Hepatic Lipid Content. Pediatr. Res. 2011, 70, 287–291. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  34. Brumbaugh, D.E.; Tearse, P.; Cree-Green, M.; Fenton, L.Z.; Brown, M.; Scherzinger, A.; Reynolds, R.; Alston, M.; Hoffman, C.; Pan, Z.; et al. Intrahepatic fat is increased in the neonatal offspring of obese women with gestational diabetes. J. Pediatr. 2013, 162, 930–936.e1. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  35. Leonard, S.A.; Rasmussen, K.M.; King, J.C.; Abrams, B. Trajectories of maternal weight from before pregnancy through postpartum and associations with childhood obesity. Am. J. Clin. Nutr. 2017, 106, 1295–1301. [Google Scholar] [CrossRef] [Green Version]
  36. Gaillard, R.; Welten, M.; Oddy, W.H.; Beilin, L.J.; Mori, T.A.; Jaddoe, V.W.; Huang, R.C. Associations of maternal prepregnancy body mass index and gestational weight gain with cardio-metabolic risk factors in adolescent offspring: A prospective cohort study. BJOG 2016, 123, 207–216. [Google Scholar] [CrossRef]
  37. Razaz, N.; Villamor, E.; Muraca, G.M.; Bonamy, A.-K.E.; Cnattingius, S. Maternal obesity and risk of cardiovascular diseases in offspring: A population-based cohort and sibling-controlled study. Lancet Diabetes Endocrinol. 2020, 8, 572–581. [Google Scholar] [CrossRef]
  38. Samuelsson, A.-M.; Matthews, P.A.; Argenton, M.; Christie, M.R.; McConnell, J.M.; Jansen, E.H.J.M.; Piersma, A.H.; Ozanne, S.E.; Twinn, D.F.; Remacle, C.; et al. Diet-Induced Obesity in Female Mice Leads to Offspring Hyperphagia, Adiposity, Hypertension, and Insulin Resistance. Hypertension 2008, 51, 383–392. [Google Scholar] [CrossRef] [Green Version]
  39. Gaillard, R.; Steegers, E.A.P.; Franco, O.H.; Hofman, A.; Jaddoe, V.W.V. Maternal weight gain in different periods of pregnancy and childhood cardio-metabolic outcomes. The Generation R Study. Int. J. Obes. 2015, 39, 677–685. [Google Scholar] [CrossRef]
  40. Kaar, J.L.; Crume, T.; Brinton, J.T.; Bischoff, K.J.; McDuffie, R.; Dabelea, D. Maternal obesity, gestational weight gain, and offspring adiposity: The exploring perinatal outcomes among children study. J. Pediatr. 2014, 165, 509–515. [Google Scholar] [CrossRef] [Green Version]
  41. Bellatorre, A.; Scherzinger, A.; Stamm, E.; Martinez, M.; Ringham, B.; Dabelea, D. Fetal Overnutrition and Adolescent Hepatic Fat Fraction: The Exploring Perinatal Outcomes in Children Study. J. Pediatr. 2018, 192, 165–170.e1. [Google Scholar] [CrossRef]
  42. Sanchez, C.E.; Barry, C.; Sabhlok, A.; Russell, K.; Majors, A.; Kollins, S.H.; Fuemmeler, B.F. Maternal pre-pregnancy obesity and child neurodevelopmental outcomes: A meta-analysis. Obes. Rev. 2018, 19, 464–484. [Google Scholar] [CrossRef] [PubMed]
  43. Razaz, N.; Tedroff, K.; Villamor, E.; Cnattingius, S. Maternal Body Mass Index in Early Pregnancy and Risk of Epilepsy in Offspring. JAMA Neurol. 2017, 74, 668–676. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  44. Li, C.; Kaur, H.; Choi, W.S.; Huang, T.T.; Lee, R.E.; Ahluwalia, J.S. Additive interactions of maternal prepregnancy BMI and breast-feeding on childhood overweight. Obes. Res. 2005, 13, 362–371. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  45. De Paula Simino, L.A.; de Fante, T.; Figueiredo Fontana, M.; Oliveira Borges, F.; Torsoni, M.A.; Milanski, M.; Velloso, L.A.; Souza Torsoni, A. Lipid overload during gestation and lactation can independently alter lipid homeostasis in offspring and promote metabolic impairment after new challenge to high-fat diet. Nutr. Metab. 2017, 14, 16. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  46. Mitanchez, D.; Jacqueminet, S.; Nizard, J.; Tanguy, M.L.; Ciangura, C.; Lacorte, J.M.; De Carne, C.; Foix L’Helias, L.; Chavatte-Palmer, P.; Charles, M.A.; et al. Effect of maternal obesity on birthweight and neonatal fat mass: A prospective clinical trial. PLoS ONE 2017, 12, e0181307. [Google Scholar] [CrossRef] [PubMed]
  47. Andres, A.; Hull, H.R.; Shankar, K.; Casey, P.H.; Cleves, M.A.; Badger, T.M. Longitudinal body composition of children born to mothers with normal weight, overweight, and obesity. Obesity 2015, 23, 1252–1258. [Google Scholar] [CrossRef]
  48. Eriksson, J.G.; Sandboge, S.; Salonen, M.K.; Kajantie, E.; Osmond, C. Long-term consequences of maternal overweight in pregnancy on offspring later health: Findings from the Helsinki Birth Cohort Study. Ann. Med. 2014, 46, 434–438. [Google Scholar] [CrossRef]
  49. Ayonrinde, O.T.; Adams, L.A.; Mori, T.A.; Beilin, L.J.; de Klerk, N.; Pennell, C.E.; White, S.; Olynyk, J.K. Sex differences between parental pregnancy characteristics and nonalcoholic fatty liver disease in adolescents. Hepatology 2018, 67, 108–122. [Google Scholar] [CrossRef] [Green Version]
  50. Deardorff, J.; Smith, L.H.; Petito, L.; Kim, H.; Abrams, B.F. Maternal Prepregnancy Weight and Children’s Behavioral and Emotional Outcomes. Am. J. Prev. Med. 2017, 53, 432–440. [Google Scholar] [CrossRef] [Green Version]
  51. Buyken, A.E.; Karaolis-Danckert, N.; Remer, T.; Bolzenius, K.; Landsberg, B.; Kroke, A. Effects of Breastfeeding on Trajectories of Body Fat and BMI throughout Childhood. Obesity 2008, 16, 389–395. [Google Scholar] [CrossRef]
  52. Gaillard, R.; Steegers, E.A.P.; Duijts, L.; Felix, J.F.; Hofman, A.; Franco, O.H.; Jaddoe, V.W.V. Childhood Cardiometabolic Outcomes of Maternal Obesity During Pregnancy. Hypertension 2014, 63, 683–691. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  53. Herrera, E.; Amusquivar, E.; Lopez-Soldado, I.; Ortega, H. Maternal lipid metabolism and placental lipid transfer. Horm. Res. 2006, 65 (Suppl. 3), 59–64. [Google Scholar] [CrossRef] [PubMed]
  54. Casart Quintero, Y.; Garrido Cisneros, D.; Guevara Flores, C.; Castillo Andrade, R.; Salas Salas, H.; Hernández Guerra, H. Perfil lipídico en embarazadas durante el tercer trimestre según índice de masa corporal y consumo de grasas. Rev. Cuba. De Obstet. Y Ginecol. 2016, 42. Available online: http://scielo.sld.cu/scielo.php?pid=S0138-600X2016000100006&script=sci_arttext&tlng=en (accessed on 4 November 2020).
  55. Herrera, E.; Ortega-Senovilla, H. Lipid metabolism during pregnancy and its implications for fetal growth. Curr. Pharm. Biotechnol. 2014, 15, 24–31. [Google Scholar] [CrossRef]
  56. Lewis, R.M.; Wadsack, C.; Desoye, G. Placental fatty acid transfer. Curr. Opin. Clin. Nutr. Metab. Care 2018, 21, 78–82. [Google Scholar] [CrossRef]
  57. Coustan, D.R. Maternal Metabolic Adaptation to Pregnancy. In Gestational Diabetes. A Decade after the HAPO Study; Lapolla, A., Metzger, B.E., Eds.; Karger Press: Basel, Switzerland, 2020; Volume 28, pp. 11–20. [Google Scholar]
  58. Childs, C.E.; Hoile, S.P.; Burdge, G.C.; Calder, P.C. Changes in rat n-3 and n-6 fatty acid composition during pregnancy are associated with progesterone concentrations and hepatic FADS2 expression. Prostaglandins Leukot. Essent. Fat. Acids 2012, 86, 141–147. [Google Scholar] [CrossRef]
  59. Stewart, F.; Rodie, V.A.; Ramsay, J.E.; Greer, I.A.; Freeman, D.J.; Meyer, B.J. Longitudinal assessment of erythrocyte fatty acid composition throughout pregnancy and post partum. Lipids 2007, 42, 335–344. [Google Scholar] [CrossRef]
  60. Scifres, C.M.; Catov, J.M.; Simhan, H.N. The impact of maternal obesity and gestational weight gain on early and mid-pregnancy lipid profiles. Obesity. 2014, 22, 932–938. [Google Scholar] [CrossRef] [Green Version]
  61. Gil-Sanchez, A.; Demmelmair, H.; Parrilla, J.J.; Koletzko, B.; Larque, E. Mechanisms involved in the selective transfer of long chain polyunsaturated Fatty acids to the fetus. Front. Genet. 2011, 2, 57. [Google Scholar] [CrossRef] [Green Version]
  62. Gil-Sanchez, A.; Larque, E.; Demmelmair, H.; Acien, M.I.; Faber, F.L.; Parrilla, J.J.; Koletzko, B. Maternal-fetal in vivo transfer of [13C]docosahexaenoic and other fatty acids across the human placenta 12 h after maternal oral intake. Am. J. Clin. Nutr. 2010, 92, 115–122. [Google Scholar] [CrossRef] [Green Version]
  63. Larque, E.; Demmelmair, H.; Berger, B.; Hasbargen, U.; Koletzko, B. In vivo investigation of the placental transfer of (13)C-labeled fatty acids in humans. J. Lipid Res. 2003, 44, 49–55. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  64. Kuipers, R.S.; Luxwolda, M.F.; Sango, W.S.; Kwesigabo, G.; Dijck-Brouwer, D.A.J.; Muskiet, F.A.J. Maternal DHA Equilibrium during Pregnancy and Lactation Is Reached at an Erythrocyte DHA Content of 8 g/100 g Fatty Acids. J. Nutr. 2011, 141, 418–427. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  65. Larque, E.; Demmelmair, H.; Gil-Sanchez, A.; Prieto-Sanchez, M.T.; Blanco, J.E.; Pagan, A.; Faber, F.L.; Zamora, S.; Parrilla, J.J.; Koletzko, B. Placental transfer of fatty acids and fetal implications. Am. J. Clin. Nutr. 2011, 94, 1908S–1913S. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  66. Food and Agriculture Organization of the United Nations. Fats and fatty acids in human nutrition. Report of an expert consultation. FAO Food. Nutr. Pap. 2010, 91, 1–166. [Google Scholar]
  67. Coletta, J.M.; Bell, S.J.; Roman, A.S. Omega-3 Fatty acids and pregnancy. Rev. Obstet. Gynecol. 2010, 3, 163–171. [Google Scholar]
  68. Koletzko, B.; Lien, E.; Agostoni, C.; Bohles, H.; Campoy, C.; Cetin, I.; Decsi, T.; Dudenhausen, J.W.; Dupont, C.; Forsyth, S.; et al. The roles of long-chain polyunsaturated fatty acids in pregnancy, lactation and infancy: Review of current knowledge and consensus recommendations. J. Perinat. Med. 2008, 36, 5–14. [Google Scholar] [CrossRef] [Green Version]
  69. Trumbo, P.; Schlicker, S.; Yates, A.A.; Poos, M.; Food and Nutrition Board of the Institute of Medicine; The National Academies. Dietary reference intakes for energy, carbohydrate, fiber, fat, fatty acids, cholesterol, protein and amino acids. J. Am. Diet Assoc. 2002, 102, 1621–1630. [Google Scholar] [CrossRef]
  70. Aranceta, J.; Pérez-Rodrigo, C. Recommended dietary reference intakes, nutritional goals and dietary guidelines for fat and fatty acids: A systematic review. Br. J. Nutr. 2012, 107 (Suppl. 2), S8–S22. [Google Scholar] [CrossRef]
  71. Institute of Medicine Committee on Strategies to Reduce Sodium Intake. The National Academies Collection: Reports funded by National Institutes of Health. In Weight Gain During Pregnancy: Reexamining the Guidelines; Rasmussen, K.M., Yaktine, A.L., Eds.; National Academies Press (US), National Academy of Sciences: Washington, DC, USA, 2009. [Google Scholar] [CrossRef]
  72. Koletzko, B.; Cetin, I.; Brenna, J.T.; Perinatal Lipid Intake Working Group; Child Health Foundation; Diabetic Pregnancy Study Group; European Association of Perinatal Medicine; European Association of Perinatal Medicine; European Society for Clinical Nutrition Metabolism; European Society for Paediatric Gastroenterology Hepatology Nutrition Committee on Nutrition; et al. Dietary fat intakes for pregnant and lactating women. Br. J. Nutr. 2007, 98, 873–877. [Google Scholar] [CrossRef] [Green Version]
  73. Jensen, C.L. Effects of n-3 fatty acids during pregnancy and lactation. Am. J. Clin. Nutr. 2006, 83, 1452S–1457S. [Google Scholar] [CrossRef] [Green Version]
  74. Much, D.; Brunner, S.; Vollhardt, C.; Schmid, D.; Sedlmeier, E.M.; Bruderl, M.; Heimberg, E.; Bartke, N.; Boehm, G.; Bader, B.L.; et al. Effect of dietary intervention to reduce the n-6/n-3 fatty acid ratio on maternal and fetal fatty acid profile and its relation to offspring growth and body composition at 1 year of age. Eur. J. Clin. Nutr. 2013, 67, 282–288. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  75. Bergmann, R.L.; Bergmann, K.E.; Haschke-Becher, E.; Richter, R.; Dudenhausen, J.W.; Barclay, D.; Haschke, F. Does maternal docosahexaenoic acid supplementation during pregnancy and lactation lower BMI in late infancy? J. Perinat. Med. 2007, 35, 295–300. [Google Scholar] [CrossRef] [PubMed]
  76. Bergmann, R.L.; Karl, E.B.; Rolf, R.; Elisabeth, H.-B.; Wolfgang, H.; Joachim, W.D. Does docosahexaenoic acid (DHA) status in pregnancy have any impact on postnatal growth? Six-year follow-up of a prospective randomized double-blind monocenter study on low-dose DHA supplements. J. Perinat. Med. 2012, 40, 677–684. [Google Scholar] [CrossRef]
  77. Bozkurt, L.; Gobl, C.S.; Hormayer, A.T.; Luger, A.; Pacini, G.; Kautzky-Willer, A. The impact of preconceptional obesity on trajectories of maternal lipids during gestation. Sci. Rep. 2016, 6, 29971. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  78. Roland, M.C.P.; Lekva, T.; Godang, K.; Bollerslev, J.; Henriksen, T. Changes in maternal blood glucose and lipid concentrations during pregnancy differ by maternal body mass index and are related to birthweight: A prospective, longitudinal study of healthy pregnancies. PLoS ONE 2020, 15, e0232749. [Google Scholar] [CrossRef]
  79. Merzouk, H.; Meghelli-Bouchenak, M.; Loukidi, B.; Prost, J.; Belleville, J. Impaired serum lipids and lipoproteins in fetal macrosomia related to maternal obesity. Biol. Neonate 2000, 77, 17–24. [Google Scholar] [CrossRef]
  80. Geraghty, A.A.; Alberdi, G.; O’Sullivan, E.J.; O’Brien, E.C.; Crosbie, B.; Twomey, P.J.; McAuliffe, F.M. Maternal and fetal blood lipid concentrations during pregnancy differ by maternal body mass index: Findings from the ROLO study. BMC Pregnancy Childbirth 2017, 17, 360. [Google Scholar] [CrossRef]
  81. Vidakovic, A.J.; Jaddoe, V.W.V.; Gishti, O.; Felix, J.F.; Williams, M.A.; Hofman, A.; Demmelmair, H.; Koletzko, B.; Tiemeier, H.; Gaillard, R. Body mass index, gestational weight gain and fatty acid concentrations during pregnancy: The Generation R Study. Eur. J. Epidemiol. 2015, 30, 1175–1185. [Google Scholar] [CrossRef] [Green Version]
  82. Cinelli, G.; Fabrizi, M.; Ravà, L.; Ciofi Degli Atti, M.; Vernocchi, P.; Vallone, C.; Pietrantoni, E.; Lanciotti, R.; Signore, F.; Manco, M. Influence of Maternal Obesity and Gestational Weight Gain on Maternal and Foetal Lipid Profile. Nutrients 2016, 8, 368. [Google Scholar] [CrossRef] [Green Version]
  83. Hellmuth, C.; Lindsay, K.L.; Uhl, O.; Buss, C.; Wadhwa, P.D.; Koletzko, B.; Entringer, S. Association of maternal prepregnancy BMI with metabolomic profile across gestation. Int. J. Obes. 2017, 41, 159–169. [Google Scholar] [CrossRef] [Green Version]
  84. Azzaroli, F.; Mazzella, G.; Marchesini, G.; Brodosi, L.; Petroni, M.L. Fatty liver in pregnancy: A narrative review of two distinct conditions. Expert Rev. Gastroenterol. Hepatol. 2020, 14, 127–135. [Google Scholar] [CrossRef] [PubMed]
  85. Araya, J.; Rodrigo, R.; Pettinelli, P.; Araya, A.V.; Poniachik, J.; Videla, L.A. Decreased liver fatty acid delta-6 and delta-5 desaturase activity in obese patients. Obesity 2010, 18, 1460–1463. [Google Scholar] [CrossRef] [PubMed]
  86. Valenzuela, R.; Echeverria, F.; Ortiz, M.; Rincón-Cervera, M.Á.; Espinosa, A.; Hernandez-Rodas, M.C.; Illesca, P.; Valenzuela, A.; Videla, L.A. Hydroxytyrosol prevents reduction in liver activity of Δ-5 and Δ-6 desaturases, oxidative stress, and depletion in long chain polyunsaturated fatty acid content in different tissues of high-fat diet fed mice. Lipids Health Dis. 2017, 16, 64. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  87. Rincón-Cervera, M.A.; Valenzuela, R.; Hernandez-Rodas, M.C.; Marambio, M.; Espinosa, A.; Mayer, S.; Romero, N.; Barrera, M.S.C.; Valenzuela, A.; Videla, L.A. Supplementation with antioxidant-rich extra virgin olive oil prevents hepatic oxidative stress and reduction of desaturation capacity in mice fed a high-fat diet: Effects on fatty acid composition in liver and extrahepatic tissues. Nutrition 2016, 32, 1254–1267. [Google Scholar] [CrossRef] [PubMed]
  88. Thompson, L.P.; Al-Hasan, Y. Impact of oxidative stress in fetal programming. J. Pregnancy 2012, 2012, 582748. [Google Scholar] [CrossRef] [PubMed]
  89. Avila, J.G.; Echeverri, I.; de Plata, C.A.; Castillo, A. Impact of oxidative stress during pregnancy on fetal epigenetic patterns and early origin of vascular diseases. Nutr. Rev. 2015, 73, 12–21. [Google Scholar] [CrossRef] [PubMed]
  90. Vega, C.C.; Reyes-Castro, L.A.; Bautista, C.J.; Larrea, F.; Nathanielsz, P.W.; Zambrano, E. Exercise in obese female rats has beneficial effects on maternal and male and female offspring metabolism. Int. J. Obes. 2015, 39, 712–719. [Google Scholar] [CrossRef] [Green Version]
  91. Leghi, G.E.; Muhlhausler, B.S. The effect of n-3 LCPUFA supplementation on oxidative stress and inflammation in the placenta and maternal plasma during pregnancy. Prostaglandins Leukot. Essent. Fat. Acids 2016, 113, 33–39. [Google Scholar] [CrossRef]
  92. Guttmacher, A.; Maddox, Y.; Spong, C. The Human Placenta Project: Placental Structure, Development, and Function in Real Time. Placenta 2014, 35, 303–304. [Google Scholar] [CrossRef] [Green Version]
  93. Gil-Sanchez, A.; Koletzko, B.; Larque, E. Current understanding of placental fatty acid transport. Curr. Opin. Clin. Nutr. Metab. Care 2012, 15, 265–272. [Google Scholar] [CrossRef]
  94. Lager, S.; Powell, T.L. Regulation of nutrient transport across the placenta. J. Pregnancy 2012, 2012, 179827. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  95. Perazzolo, S.; Hirschmugl, B.; Wadsack, C.; Desoye, G.; Lewis, R.M.; Sengers, B.G. The influence of placental metabolism on fatty acid transfer to the fetus. J. Lipid Res. 2017, 58, 443–454. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  96. Duttaroy, A.K. Transport of fatty acids across the human placenta: A review. Prog. Lipid Res. 2009, 48, 52–61. [Google Scholar] [CrossRef] [PubMed]
  97. Prieto-Sánchez, M.T.; Ruiz-Palacios, M.; Blanco-Carnero, J.E.; Pagan, A.; Hellmuth, C.; Uhl, O.; Peissner, W.; Ruiz-Alcaraz, A.J.; Parrilla, J.J.; Koletzko, B.; et al. Placental MFSD2a transporter is related to decreased DHA in cord blood of women with treated gestational diabetes. Clin. Nutr. 2017, 36, 513–521. [Google Scholar] [CrossRef] [PubMed]
  98. Nguyen, L.N.; Ma, D.; Shui, G.; Wong, P.; Cazenave-Gassiot, A.; Zhang, X.; Wenk, M.R.; Goh, E.L.K.; Silver, D.L. Mfsd2a is a transporter for the essential omega-3 fatty acid docosahexaenoic acid. Nature 2014, 509, 503–506. [Google Scholar] [CrossRef]
  99. Roberts, K.A.; Riley, S.C.; Reynolds, R.M.; Barr, S.; Evans, M.; Statham, A.; Hor, K.; Jabbour, H.N.; Norman, J.E.; Denison, F.C. Placental structure and inflammation in pregnancies associated with obesity. Placenta 2011, 32, 247–254. [Google Scholar] [CrossRef]
  100. Lager, S.; Jansson, N.; Olsson, A.L.; Wennergren, M.; Jansson, T.; Powell, T.L. Effect of IL-6 and TNF-alpha on fatty acid uptake in cultured human primary trophoblast cells. Placenta 2011, 32, 121–127. [Google Scholar] [CrossRef]
  101. Lappas, M.; Hiden, U.; Desoye, G.; Froehlich, J.; Hauguel-de Mouzon, S.; Jawerbaum, A. The role of oxidative stress in the pathophysiology of gestational diabetes mellitus. Antioxid. Redox Signal. 2011, 15, 3061–3100. [Google Scholar] [CrossRef]
  102. Jones, M.L.; Mark, P.J.; Keelan, J.A.; Barden, A.; Mas, E.; Mori, T.A.; Waddell, B.J. Maternal dietary omega-3 fatty acid intake increases resolvin and protectin levels in the rat placenta. J. Lipid Res. 2013, 54, 2247–2254. [Google Scholar] [CrossRef] [Green Version]
  103. Keelan, J.A.; Mas, E.; D’Vaz, N.; Dunstan, J.A.; Li, S.; Barden, A.E.; Mark, P.J.; Waddell, B.J.; Prescott, S.L.; Mori, T.A. Effects of maternal n-3 fatty acid supplementation on placental cytokines, pro-resolving lipid mediators and their precursors. Reproduction 2015, 149, 171. [Google Scholar] [CrossRef] [Green Version]
  104. Jones, M.L.; Mark, P.J.; Mori, T.A.; Keelan, J.A.; Waddell, B.J. Maternal Dietary Omega-3 Fatty Acid Supplementation Reduces Placental Oxidative Stress and Increases Fetal and Placental Growth in the Rat1. Biol. Reprod. 2013, 88. [Google Scholar] [CrossRef]
  105. Jones, M.; Mark, P.; Waddell, B. Maternal dietary omega-3 fatty acids and placental function. Reproduction 2014, 147, R143–R152. [Google Scholar] [CrossRef] [Green Version]
  106. Shrestha, N.; Holland, O.J.; Kent, N.L.; Perkins, A.V.; McAinch, A.J.; Cuffe, J.S.M.; Hryciw, D.H. Maternal High Linoleic Acid Alters Placental Fatty Acid Composition. Nutrients 2020, 12, 2183. [Google Scholar] [CrossRef] [PubMed]
  107. Reyes-Hernández, C.G.; Ramiro-Cortijo, D.; Rodríguez-Rodríguez, P.; Giambelluca, S.; Simonato, M.; González, M.D.C.; López de Pablo, A.L.; López-Giménez, M.D.R.; Cogo, P.; Sáenz de Pipaón, M.; et al. Effects of Arachidonic and Docosohexahenoic Acid Supplementation during Gestation in Rats. Implication of Placental Oxidative Stress. Int. J. Mol. Sci. 2018, 19, 3863. [Google Scholar] [CrossRef] [Green Version]
  108. Garcia-Contreras, C.; Vazquez-Gomez, M.; Astiz, S.; Torres-Rovira, L.; Sanchez-Sanchez, R.; Gomez-Fidalgo, E.; Gonzalez, J.; Isabel, B.; Rey, A.; Ovilo, C.; et al. Ontogeny of Sex-Related Differences in Foetal Developmental Features, Lipid Availability and Fatty Acid Composition. Int. J. Mol. Sci. 2017, 18, 1171. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  109. Zhu, M.J.; Ma, Y.; Long, N.M.; Du, M.; Ford, S.P. Maternal obesity markedly increases placental fatty acid transporter expression and fetal blood triglycerides at midgestation in the ewe. Am. J. Physiol. Regul. Integr. Comp. Physiol. 2010, 299, R1224–R1231. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  110. Hirschmugl, B.; Desoye, G.; Catalano, P.; Klymiuk, I.; Scharnagl, H.; Payr, S.; Kitzinger, E.; Schliefsteiner, C.; Lang, U.; Wadsack, C.; et al. Maternal obesity modulates intracellular lipid turnover in the human term placenta. Int. J. Obes. 2017, 41, 317–323. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  111. Dube, E.; Gravel, A.; Martin, C.; Desparois, G.; Moussa, I.; Ethier-Chiasson, M.; Forest, J.C.; Giguere, Y.; Masse, A.; Lafond, J. Modulation of fatty acid transport and metabolism by maternal obesity in the human full-term placenta. Biol. Reprod. 2012, 87, 14, 1–11. [Google Scholar] [CrossRef]
  112. Lager, S.; Ramirez, V.I.; Gaccioli, F.; Jang, B.; Jansson, T.; Powell, T.L. Protein expression of fatty acid transporter 2 is polarized to the trophoblast basal plasma membrane and increased in placentas from overweight/obese women. Placenta 2016, 40, 60–66. [Google Scholar] [CrossRef] [Green Version]
  113. Calabuig-Navarro, V.; Haghiac, M.; Minium, J.; Glazebrook, P.; Ranasinghe, G.C.; Hoppel, C.; Hauguel de-Mouzon, S.; Catalano, P.; O’Tierney-Ginn, P. Effect of Maternal Obesity on Placental Lipid Metabolism. Endocrinology 2017, 158, 2543–2555. [Google Scholar] [CrossRef]
  114. Draycott, S.A.V.; Liu, G.; Daniel, Z.C.; Elmes, M.J.; Muhlhausler, B.S.; Langley-Evans, S.C. Maternal dietary ratio of linoleic acid to alpha-linolenic acid during pregnancy has sex-specific effects on placental and fetal weights in the rat. Nutr. Metab. 2019, 16, 1. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  115. Fattuoni, C.; Mando, C.; Palmas, F.; Anelli, G.M.; Novielli, C.; Parejo Laudicina, E.; Savasi, V.M.; Barberini, L.; Dessi, A.; Pintus, R.; et al. Preliminary metabolomics analysis of placenta in maternal obesity. Placenta 2018, 61, 89–95. [Google Scholar] [CrossRef] [PubMed]
  116. Brass, E.; Hanson, E.; O’Tierney-Ginn, P.F. Placental oleic acid uptake is lower in male offspring of obese women. Placenta 2013, 34, 503–509. [Google Scholar] [CrossRef] [Green Version]
  117. Gallardo, J.M.; Gomez-Lopez, J.; Medina-Bravo, P.; Juarez-Sanchez, F.; Contreras-Ramos, A.; Galicia-Esquivel, M.; Sanchez-Urbina, R.; Klunder-Klunder, M. Maternal obesity increases oxidative stress in the newborn. Obesity 2015, 23, 1650–1654. [Google Scholar] [CrossRef] [PubMed]
  118. Shoji, H.; Franke, C.; Demmelmair, H.; Koletzko, B. Effect of docosahexaenoic acid on oxidative stress in placental trophoblast cells. Early Hum. Dev. 2009, 85, 433–437. [Google Scholar] [CrossRef] [PubMed]
  119. Jansson, T.; Powell, T.L. Role of the placenta in fetal programming: Underlying mechanisms and potential interventional approaches. Clin. Sci. 2007, 113, 1–13. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  120. Marciniak, A.; Patro-Malysza, J.; Kimber-Trojnar, Z.; Marciniak, B.; Oleszczuk, J.; Leszczynska-Gorzelak, B. Fetal programming of the metabolic syndrome. Taiwan J. Obstet. Gynecol 2017, 56, 133–138. [Google Scholar] [CrossRef]
  121. Benassayag, C.; Mignot, T.M.; Haourigui, M.; Civel, C.; Hassid, J.; Carbonne, B.; Nunez, E.A.; Ferre, F. High polyunsaturated fatty acid, thromboxane A2, and alpha-fetoprotein concentrations at the human feto-maternal interface. J. Lipid Res. 1997, 38, 276–286. [Google Scholar]
  122. Su, H.M.; Corso, T.N.; Nathanielsz, P.W.; Brenna, J.T. Linoleic acid kinetics and conversion to arachidonic acid in the pregnant and fetal baboon. J. Lipid Res. 1999, 40, 1304–1312. [Google Scholar]
  123. Chambaz, J.; Ravel, D.; Manier, M.C.; Pepin, D.; Mulliez, N.; Bereziat, G. Essential Fatty Acids Interconversion in the Human Fetal Liver. Neonatology 1985, 47, 136–140. [Google Scholar] [CrossRef]
  124. Rodriguez, A.; Sarda, P.; Nessmann, C.; Boulot, P.; Leger, C.L.; Descomps, B. Δ6- and Δ5-desaturase activities in the human fetal liver: Kinetic aspects. J. Lipid Res. 1998, 39, 1825–1832. [Google Scholar] [PubMed]
  125. Haggarty, P. Fatty Acid Supply to the Human Fetus. Annu. Rev. Nutr. 2010, 30, 237–255. [Google Scholar] [CrossRef] [PubMed]
  126. Wesolowski, S.R.; Kasmi, K.C.; Jonscher, K.R.; Friedman, J.E. Developmental origins of NAFLD: A womb with a clue. Nat. Rev. Gastroenterol. Hepatol. 2017, 14, 81–96. [Google Scholar] [CrossRef] [Green Version]
  127. Villee, C.A.; Hagerman, D.D. Effect of oxygen deprivation on the metabolism of fetal and adult tissues. Am. J. Physiol. 1958, 194, 457–464. [Google Scholar] [CrossRef] [PubMed]
  128. Saggerson, E.D.; Carpenter, C.A. Regulation of hepatic carnitine palmitoyltransferase activity during the foetal-neonatal transition. FEBS Lett. 1982, 150, 177–180. [Google Scholar] [CrossRef] [Green Version]
  129. Oey, N.A.; Ruiter, J.P.; Attie-Bitach, T.; Ijlst, L.; Wanders, R.J.; Wijburg, F.A. Fatty acid oxidation in the human fetus: Implications for fetal and adult disease. J. Inherit. Metab. Dis. 2006, 29, 71–75. [Google Scholar] [CrossRef]
  130. Diaz, P.; Harris, J.; Rosario, F.J.; Powell, T.L.; Jansson, T. Increased placental fatty acid transporter 6 and binding protein 3 expression and fetal liver lipid accumulation in a mouse model of obesity in pregnancy. Am. J. Physiol. Regul. Integr. Comp. Physiol. 2015, 309, R1569–R1577. [Google Scholar] [CrossRef] [Green Version]
  131. McCurdy, C.E.; Bishop, J.M.; Williams, S.M.; Grayson, B.E.; Smith, M.S.; Friedman, J.E.; Grove, K.L. Maternal high-fat diet triggers lipotoxicity in the fetal livers of nonhuman primates. J. Clin. Investig. 2009, 119, 323–335. [Google Scholar] [CrossRef] [Green Version]
  132. Mazzucco, M.B.; Fornes, D.; Capobianco, E.; Higa, R.; Jawerbaum, A.; White, V. Maternal saturated-fat-rich diet promotes leptin resistance in fetal liver lipid catabolism and programs lipid homeostasis impairments in the liver of rat offspring. J. Nutr. Biochem. 2016, 27, 61–69. [Google Scholar] [CrossRef]
  133. Cerf, M.E.; Louw, J.; Herrera, E. High Fat Diet Exposure during Fetal Life Enhances Plasma and Hepatic Omega-6 Fatty Acid Profiles in Fetal Wistar Rats. Nutrients 2015, 7, 7231–7241. [Google Scholar] [CrossRef] [Green Version]
  134. Ramaiyan, B.; Talahalli, R.R. Dietary Unsaturated Fatty Acids Modulate Maternal Dyslipidemia-Induced DNA Methylation and Histone Acetylation in Placenta and Fetal Liver in Rats. Lipids 2018, 53, 581–588. [Google Scholar] [CrossRef] [PubMed]
  135. Novak, E.M.; Keller, B.O.; Innis, S.M. Metabolic development in the liver and the implications of the n-3 fatty acid supply. Am. J. Physiol. Gastrointest. Liver Physiol. 2012, 302, G250–G259. [Google Scholar] [CrossRef] [Green Version]
  136. Serafim, T.L.; Cunha-Oliveira, T.; Deus, C.M.; Sardao, V.A.; Cardoso, I.M.; Yang, S.; Odhiambo, J.F.; Ghnenis, A.B.; Smith, A.M.; Li, J.; et al. Maternal obesity in sheep impairs fetal hepatic respiratory chain capacity. Eur. J. Clin. Investig. 2020, e13375. [Google Scholar] [CrossRef]
  137. Stewart, M.S.; Heerwagen, M.J.; Friedman, J.E. Developmental programming of pediatric nonalcoholic fatty liver disease: Redefining the “first hit”. Clin. Obstet. Gynecol. 2013, 56, 577–590. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  138. Baker, P.R., 2nd; Friedman, J.E. Mitochondrial role in the neonatal predisposition to developing nonalcoholic fatty liver disease. J. Clin. Investig. 2018, 128, 3692–3703. [Google Scholar] [CrossRef] [PubMed]
  139. Bobiński, R.; Mikulska, M. The ins and outs of maternal-fetal fatty acid metabolism. Acta Biochim. Pol. 2015, 62. [Google Scholar] [CrossRef]
  140. Ailhaud, G.; Massiera, F.; Weill, P.; Legrand, P.; Alessandri, J.M.; Guesnet, P. Temporal changes in dietary fats: Role of n-6 polyunsaturated fatty acids in excessive adipose tissue development and relationship to obesity. Prog. Lipid Res. 2006, 45, 203–236. [Google Scholar] [CrossRef]
  141. Yan, X.; Zhu, M.J.; Xu, W.; Tong, J.F.; Ford, S.P.; Nathanielsz, P.W.; Du, M. Up-regulation of Toll-like receptor 4/nuclear factor-kappaB signaling is associated with enhanced adipogenesis and insulin resistance in fetal skeletal muscle of obese sheep at late gestation. Endocrinology 2010, 151, 380–387. [Google Scholar] [CrossRef]
  142. Zhu, M.J.; Han, B.; Tong, J.; Ma, C.; Kimzey, J.M.; Underwood, K.R.; Xiao, Y.; Hess, B.W.; Ford, S.P.; Nathanielsz, P.W.; et al. AMP-activated protein kinase signalling pathways are down regulated and skeletal muscle development impaired in fetuses of obese, over-nourished sheep. J. Physiol. 2008, 586, 2651–2664. [Google Scholar] [CrossRef]
  143. Hsueh, T.Y.; Baum, J.I.; Huang, Y. Effect of Eicosapentaenoic Acid and Docosahexaenoic Acid on Myogenesis and Mitochondrial Biosynthesis during Murine Skeletal Muscle Cell Differentiation. Front. Nutr. 2018, 5, 15. [Google Scholar] [CrossRef] [Green Version]
  144. Chouinard-Watkins, R.; Lacombe, R.J.S.; Bazinet, R.P. Mechanisms regulating brain docosahexaenoic acid uptake: What is the recent evidence? Curr. Opin. Clin. Nutr. Metab. Care 2018, 21, 71–77. [Google Scholar] [CrossRef] [PubMed]
  145. Hadley, K.B.; Ryan, A.S.; Forsyth, S.; Gautier, S.; Salem, N. The Essentiality of Arachidonic Acid in Infant Development. Nutrients 2016, 8, 216. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  146. Guemez-Gamboa, A.; Nguyen, L.N.; Yang, H.; Zaki, M.S.; Kara, M.; Ben-Omran, T.; Akizu, N.; Rosti, R.O.; Rosti, B.; Scott, E.; et al. Inactivating mutations in MFSD2A, required for omega-3 fatty acid transport in brain, cause a lethal microcephaly syndrome. Nat. Genet. 2015, 47, 809–813. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  147. Seong, J.; Kang, J.Y.; Sun, J.S.; Kim, K.W. Hypothalamic inflammation and obesity: A mechanistic review. Arch. Pharm. Res. 2019, 42, 383–392. [Google Scholar] [CrossRef] [PubMed]
  148. Neri, C.; Edlow, A.G. Effects of Maternal Obesity on Fetal Programming: Molecular Approaches. Cold Spring Harb. Perspect. Med. 2015, 6, a026591. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  149. Thaler, J.P.; Yi, C.X.; Schur, E.A.; Guyenet, S.J.; Hwang, B.H.; Dietrich, M.O.; Zhao, X.; Sarruf, D.A.; Izgur, V.; Maravilla, K.R.; et al. Obesity is associated with hypothalamic injury in rodents and humans. J. Clin. Investig. 2012, 122, 153–162. [Google Scholar] [CrossRef] [Green Version]
  150. Williams, L.M. Hypothalamic dysfunction in obesity. Proc. Nutr. Soc. 2012, 71, 521–533. [Google Scholar] [CrossRef] [Green Version]
  151. Singer, K.; Lumeng, C.N. The initiation of metabolic inflammation in childhood obesity. J. Clin. Investig. 2017, 127, 65–73. [Google Scholar] [CrossRef] [Green Version]
  152. Rother, E.; Kuschewski, R.; Alcazar, M.A.A.; Oberthuer, A.; Bae-Gartz, I.; Vohlen, C.; Roth, B.; Dötsch, J. Hypothalamic JNK1 and IKKβ Activation and Impaired Early Postnatal Glucose Metabolism after Maternal Perinatal High-Fat Feeding. Endocrinology 2012, 153, 770–781. [Google Scholar] [CrossRef]
  153. Smithers, L.G.; Markrides, M.; Gibson, R.A. Human milk fatty acids from lactating mothers of preterm infants: A study revealing wide intra- and inter-individual variation. Prostaglandins Leukot Essent Fat. Acids 2010, 83, 9–13. [Google Scholar] [CrossRef]
  154. Sadovnikova, A.; Wysolmerski, J.J.; Hovey, R.C. Chapter 14-The Onset and Maintenance of Human Lactation and its Endocrine Regulation. In Maternal-Fetal and Neonatal Endocrinology; Kovacs, C.S., Deal, C.L., Eds.; Academic Press: Cambridge, MA, USA, 2020; pp. 189–205. [Google Scholar] [CrossRef]
  155. Mohammad, M.A.; Haymond, M.W. Regulation of lipid synthesis genes and milk fat production in human mammary epithelial cells during secretory activation. Am. J. Physiol. Endocrinol. Metab. 2013, 305, E700–E716. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  156. Rudolph, M.C.; Russell, T.D.; Webb, P.; Neville, M.C.; Anderson, S.M. Prolactin-mediated regulation of lipid biosynthesis genes in vivo in the lactating mammary epithelial cell. Am. J. Physiol. Endocrinol. Metab. 2011, 300, E1059–E1068. [Google Scholar] [CrossRef] [Green Version]
  157. Wen, L.; Wu, Y.; Yang, Y.; Han, T.L.; Wang, W.; Fu, H.; Zheng, Y.; Shan, T.; Chen, J.; Xu, P.; et al. Gestational Diabetes Mellitus Changes the Metabolomes of Human Colostrum, Transition Milk and Mature Milk. Med. Sci. Monit. 2019, 25, 6128–6152. [Google Scholar] [CrossRef]
  158. Siziba, L.P.; Lorenz, L.; Stahl, B.; Mank, M.; Marosvölgyi, T.; Decsi, T.; Rothenbacher, D.; Genuneit, J. Changes in Human Milk Fatty Acid Composition During Lactation: The Ulm SPATZ Health Study. Nutrients 2019, 11, 2842. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  159. Koletzko, B.; Rodriguez-Palmero, M.; Demmelmair, H.; Fidler, N.; Jensen, R.; Sauerwald, T. Physiological aspects of human milk lipids. Early Hum. Dev. 2001, 65, S3–S18. [Google Scholar] [CrossRef]
  160. Smith, S.L.; Rouse, C.A. Docosahexaenoic acid and the preterm infant. Matern. Health Neonatol. Perinatol. 2017, 3, 22. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  161. Sherry, C.L.; Oliver, J.S.; Marriage, B.J. Docosahexaenoic acid supplementation in lactating women increases breast milk and plasma docosahexaenoic acid concentrations and alters infant omega 6:3 fatty acid ratio. Prostaglandins Leukot. Essent. Fat. Acids 2015, 95, 63–69. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  162. Rasmussen, K.M.; Kjolhede, C.L. Prepregnant overweight and obesity diminish the prolactin response to suckling in the first week postpartum. Pediatrics 2004, 113, e465–e471. [Google Scholar] [CrossRef] [Green Version]
  163. Nommsen-Rivers, L.A. Does Insulin Explain the Relation between Maternal Obesity and Poor Lactation Outcomes? An Overview of the Literature. Adv. Nutr. 2016, 7, 407–414. [Google Scholar] [CrossRef] [Green Version]
  164. Saben, J.L.; Bales, E.S.; Jackman, M.R.; Orlicky, D.; MacLean, P.S.; McManaman, J.L. Maternal obesity reduces milk lipid production in lactating mice by inhibiting acetyl-CoA carboxylase and impairing fatty acid synthesis. PLoS ONE 2014, 9, e98066. [Google Scholar] [CrossRef]
  165. Makela, J.; Vaarno, J.; Kaljonen, A.; Niinikoski, H.; Lagstrom, H. Maternal overweight impacts infant feeding patterns--the STEPS Study. Eur. J. Clin. Nutr. 2014, 68, 43–49. [Google Scholar] [CrossRef] [PubMed]
  166. Lima, M.S.; Perez, G.S.; Morais, G.L.; Santos, L.S.; Cordeiro, G.S.; Couto, R.D.; Deiró, T.C.B.J.; Leandro, C.G.; Barreto-Medeiros, J.M. Effects of maternal high fat intake during pregnancy and lactation on total cholesterol and adipose tissue in neonatal rats. Braz. J. Biol. 2018, 78, 615–618. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  167. Erliana, U.D.; Fly, A.D. The Function and Alteration of Immunological Properties in Human Milk of Obese Mothers. Nutrients 2019, 11, 1284. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  168. O’Reilly, J.R.; Reynolds, R.M. The risk of maternal obesity to the long-term health of the offspring. Clin. Endocrinol. 2013, 78, 9–16. [Google Scholar] [CrossRef] [PubMed]
  169. Makela, J.; Linderborg, K.; Niinikoski, H.; Yang, B.; Lagstrom, H. Breast milk fatty acid composition differs between overweight and normal weight women: The STEPS Study. Eur. J. Nutr. 2013, 52, 727–735. [Google Scholar] [CrossRef]
  170. De la Garza Puentes, A.; Martí Alemany, A.; Chisaguano, A.M.; Montes Goyanes, R.; Castellote, A.I.; Torres-Espínola, F.J.; García-Valdés, L.; Escudero-Marín, M.; Segura, M.T.; Campoy, C.; et al. The Effect of Maternal Obesity on Breast Milk Fatty Acids and Its Association with Infant Growth and Cognition—The PREOBE Follow-Up. Nutrients 2019, 11, 2154. [Google Scholar] [CrossRef] [Green Version]
  171. Antonakou, A.; Skenderi, K.P.; Chiou, A.; Anastasiou, C.A.; Bakoula, C.; Matalas, A.L. Breast milk fat concentration and fatty acid pattern during the first six months in exclusively breastfeeding Greek women. Eur. J. Nutr. 2013, 52, 963–973. [Google Scholar] [CrossRef]
  172. Ellsworth, L.; Perng, W.; Harman, E.; Das, A.; Pennathur, S.; Gregg, B. Impact of maternal overweight and obesity on milk composition and infant growth. Matern. Child. Nutr. 2020, 16, e12979. [Google Scholar] [CrossRef]
  173. García-Ravelo, S.; Díaz-Gómez, N.M.; Martín, M.V.; Dorta-Guerra, R.; Murray, M.; Escuder, D.; Rodríguez, C. Fatty Acid Composition and Eicosanoid Levels (LTE(4) and PGE(2)) of Human Milk from Normal Weight and Overweight Mothers. Breastfeed. Med. Off. J. Acad. Breastfeed. Med. 2018, 13, 702–710. [Google Scholar] [CrossRef]
  174. Panagos, P.G.; Vishwanathan, R.; Penfield-Cyr, A.; Matthan, N.R.; Shivappa, N.; Wirth, M.D.; Hebert, J.R.; Sen, S. Breastmilk from obese mothers has pro-inflammatory properties and decreased neuroprotective factors. J. Perinatol. 2016, 36, 284–290. [Google Scholar] [CrossRef] [Green Version]
  175. Storck Lindholm, E.; Strandvik, B.; Altman, D.; Moller, A.; Palme Kilander, C. Different fatty acid pattern in breast milk of obese compared to normal-weight mothers. Prostaglandins Leukot Essent. Fat. Acids 2013, 88, 211–217. [Google Scholar] [CrossRef] [PubMed]
  176. Grant, W.F.; Gillingham, M.B.; Batra, A.K.; Fewkes, N.M.; Comstock, S.M.; Takahashi, D.; Braun, T.P.; Grove, K.L.; Friedman, J.E.; Marks, D.L. Maternal high fat diet is associated with decreased plasma n-3 fatty acids and fetal hepatic apoptosis in nonhuman primates. PLoS ONE 2011, 6, e17261. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  177. Brezinova, M.; Kuda, O.; Hansikova, J.; Rombaldova, M.; Balas, L.; Bardova, K.; Durand, T.; Rossmeisl, M.; Cerna, M.; Stranak, Z.; et al. Levels of palmitic acid ester of hydroxystearic acid (PAHSA) are reduced in the breast milk of obese mothers. Biochim. Biophys. Acta Mol. Cell Biol. Lipids 2018, 1863, 126–131. [Google Scholar] [CrossRef] [PubMed]
  178. Mennitti, L.V.; Oliveira, J.L.; Morais, C.A.; Estadella, D.; Oyama, L.M.; Oller do Nascimento, C.M.; Pisani, L.P. Type of fatty acids in maternal diets during pregnancy and/or lactation and metabolic consequences of the offspring. J. Nutr. Biochem. 2015, 26, 99–111. [Google Scholar] [CrossRef] [PubMed]
  179. Fan, R.; Toney, A.M.; Jang, Y.; Ro, S.H.; Chung, S. Maternal n-3 PUFA supplementation promotes fetal brown adipose tissue development through epigenetic modifications in C57BL/6 mice. Biochim. Biophys. Acta Mol. Cell Biol. Lipids 2018, 1863, 1488–1497. [Google Scholar] [CrossRef] [PubMed]
  180. Argentato, P.P.; de Cassia Cesar, H.; Estadella, D.; Pisani, L.P. Programming mediated by fatty acids affects uncoupling protein 1 (UCP-1) in brown adipose tissue. Br. J. Nutr. 2018, 120, 619–627. [Google Scholar] [CrossRef] [PubMed]
  181. Muhlhausler, B.S.; Miljkovic, D.; Fong, L.; Xian, C.J.; Duthoit, E.; Gibson, R.A. Maternal omega-3 supplementation increases fat mass in male and female rat offspring. Front. Genet. 2011, 2, 48. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  182. Drouillet, P.; Forhan, A.; De Lauzon-Guillain, B.; Thiebaugeorges, O.; Goua, V.; Magnin, G.; Schweitzer, M.; Kaminski, M.; Ducimetiere, P.; Charles, M.A. Maternal fatty acid intake and fetal growth: Evidence for an association in overweight women. The ‘EDEN mother-child’ cohort (study of pre- and early postnatal determinants of the child’s development and health). Br. J. Nutr. 2009, 101, 583–591. [Google Scholar] [CrossRef] [Green Version]
  183. Monthe-Dreze, C.; Penfield-Cyr, A.; Smid, M.C.; Sen, S. Maternal Pre-Pregnancy Obesity Attenuates Response to Omega-3 Fatty Acids Supplementation During Pregnancy. Nutrients 2018, 10, 1908. [Google Scholar] [CrossRef] [Green Version]
  184. Much, D.; Brunner, S.; Vollhardt, C.; Schmid, D.; Sedlmeier, E.-M.; Brüderl, M.; Heimberg, E.; Bartke, N.; Boehm, G.; Bader, B.L.; et al. Breast milk fatty acid profile in relation to infant growth and body composition: Results from the INFAT study. Pediatr. Res. 2013, 74, 230–237. [Google Scholar] [CrossRef] [Green Version]
  185. Sorensen, L.B.; Dyssegaard, C.B.; Damsgaard, C.T.; Petersen, R.A.; Dalskov, S.M.; Hjorth, M.F.; Andersen, R.; Tetens, I.; Ritz, C.; Astrup, A.; et al. The effects of Nordic school meals on concentration and school performance in 8- to 11-year-old children in the OPUS School Meal Study: A cluster-randomised, controlled, cross-over trial. Br. J. Nutr. 2015, 113, 1280–1291. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  186. Richardson, A.J.; Montgomery, P. The Oxford-Durham study: A randomized, controlled trial of dietary supplementation with fatty acids in children with developmental coordination disorder. Pediatrics 2005, 115, 1360–1366. [Google Scholar] [CrossRef] [PubMed] [Green Version]
Figure 1. Maternal obesity is associated with changes in circulating fatty acid (FA) profile and placental alterations including a lipotoxic environment, a pro-inflammatory state, and elevated oxidative stress, that altogether lead to modifications in FA transporters that affect the development of fetal organs such as the brain, liver, adipose tissue, and skeletal muscle. Moreover, maternal obesity reduces the time of breastfeeding and alters the FA composition of breast milk.
Figure 1. Maternal obesity is associated with changes in circulating fatty acid (FA) profile and placental alterations including a lipotoxic environment, a pro-inflammatory state, and elevated oxidative stress, that altogether lead to modifications in FA transporters that affect the development of fetal organs such as the brain, liver, adipose tissue, and skeletal muscle. Moreover, maternal obesity reduces the time of breastfeeding and alters the FA composition of breast milk.
Nutrients 13 00019 g001
Figure 2. Impact of maternal obesity on the metabolism of n-3 and n-6 LCPUFA during pregnancy and lactation. An unhealthy diet, commonly associated with obesity, along with modifications in FA metabolism in the maternal liver, leads to low circulating levels of n-3 PUFA and elevated circulating levels of n-6 PUFA. These modifications are also present in the placenta, fetus, breast milk, and, consequently, in the child. LCPUFA, long chain polyunsaturated fatty acid; PUFA, polyunsaturated fatty acid; ALA, α–linolenic acid; AA, arachidonic acid; DHA, docosahexaenoic acid; DPA, docosapentaenoic acid; EPA, eicosapentaenoic acid; n-6/n-3 ratio, total polyunsaturated fatty acid n-6/n-3 ratio; CVD, cardiovascular disease; Delta -5 (Δ-5) desaturase; Delta-6 (Δ-6) desaturase.
Figure 2. Impact of maternal obesity on the metabolism of n-3 and n-6 LCPUFA during pregnancy and lactation. An unhealthy diet, commonly associated with obesity, along with modifications in FA metabolism in the maternal liver, leads to low circulating levels of n-3 PUFA and elevated circulating levels of n-6 PUFA. These modifications are also present in the placenta, fetus, breast milk, and, consequently, in the child. LCPUFA, long chain polyunsaturated fatty acid; PUFA, polyunsaturated fatty acid; ALA, α–linolenic acid; AA, arachidonic acid; DHA, docosahexaenoic acid; DPA, docosapentaenoic acid; EPA, eicosapentaenoic acid; n-6/n-3 ratio, total polyunsaturated fatty acid n-6/n-3 ratio; CVD, cardiovascular disease; Delta -5 (Δ-5) desaturase; Delta-6 (Δ-6) desaturase.
Nutrients 13 00019 g002
Figure 3. Schematic representation of placental fatty acid transport. Maternal triglycerides are hydrolyzed to non-esterified fatty acid (NEFA) by placental lipases (lipoprotein lipase (LPL) and endothelial lipoprotein lipase 9 (EL9). Then, the placenta can uptake and move the NEFA across the membrane through a flip-flop mechanism or with the help of binding proteins and transporters. FABP: fatty acid binding protein; FAT/CD36: fatty acid translocase; FATP: fatty acid transport protein; MFSD2a: main facilitator superfamily domain containing 2a; FABPpm: plasma membrane’s fatty acid binding protein.
Figure 3. Schematic representation of placental fatty acid transport. Maternal triglycerides are hydrolyzed to non-esterified fatty acid (NEFA) by placental lipases (lipoprotein lipase (LPL) and endothelial lipoprotein lipase 9 (EL9). Then, the placenta can uptake and move the NEFA across the membrane through a flip-flop mechanism or with the help of binding proteins and transporters. FABP: fatty acid binding protein; FAT/CD36: fatty acid translocase; FATP: fatty acid transport protein; MFSD2a: main facilitator superfamily domain containing 2a; FABPpm: plasma membrane’s fatty acid binding protein.
Nutrients 13 00019 g003
Table 1. Pregnant and lactating women’s dietary recommendations.
Table 1. Pregnant and lactating women’s dietary recommendations.
Fat IntakeFAO/WHO [66]Dietary Reference Intake [70,71]
Fat 20–35% E20–35% E
MUFA Difference--
SFA Max 10% E<10% E
PUFA6–10% E--
n-6 PUFA2.5–9% E5–10% E
n-3 PUFA0.5–2% E0.6–1.2% E
DHA200 mg/d1.4 g/d ALA
EPA + DHA300 mg/d--
AA800 mg/d13 g/d LA
Trans fatty acids<1% E--
% E (% of total energy); MUFA, monounsaturated fatty acids; SFA, Saturated fatty acids; PUFA, polyunsaturated fatty acids; ALA, α–linolenic acid; AA, arachidonic acid; DHA, docosahexaenoic acid; EPA, eicosapentaenoic acid; LA, linoleic acid.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Álvarez, D.; Muñoz, Y.; Ortiz, M.; Maliqueo, M.; Chouinard-Watkins, R.; Valenzuela, R. Impact of Maternal Obesity on the Metabolism and Bioavailability of Polyunsaturated Fatty Acids during Pregnancy and Breastfeeding. Nutrients 2021, 13, 19. https://doi.org/10.3390/nu13010019

AMA Style

Álvarez D, Muñoz Y, Ortiz M, Maliqueo M, Chouinard-Watkins R, Valenzuela R. Impact of Maternal Obesity on the Metabolism and Bioavailability of Polyunsaturated Fatty Acids during Pregnancy and Breastfeeding. Nutrients. 2021; 13(1):19. https://doi.org/10.3390/nu13010019

Chicago/Turabian Style

Álvarez, Daniela, Yasna Muñoz, Macarena Ortiz, Manuel Maliqueo, Raphaël Chouinard-Watkins, and Rodrigo Valenzuela. 2021. "Impact of Maternal Obesity on the Metabolism and Bioavailability of Polyunsaturated Fatty Acids during Pregnancy and Breastfeeding" Nutrients 13, no. 1: 19. https://doi.org/10.3390/nu13010019

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop