Next Article in Journal
Do Diet and Lifestyles Play a Role in the Pathogenesis of NMSCs?
Next Article in Special Issue
Sex-Specific Effects of Chronic Creatine Supplementation on Hippocampal-Mediated Spatial Cognition in the 3xTg Mouse Model of Alzheimer’s Disease
Previous Article in Journal
Maternal Linoleic Acid Overconsumption Alters Offspring Gut and Adipose Tissue Homeostasis in Young but Not Older Adult Rats
Previous Article in Special Issue
The Different Relationship between Homocysteine and Uric Acid Levels with Respect to the MTHFR C677T Polymorphism According to Gender in Patients with Cognitive Impairment
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Vitamins in Alzheimer’s Disease—Review of the Latest Reports

by
Anita Mielech
,
Anna Puścion-Jakubik
*,†,
Renata Markiewicz-Żukowska
and
Katarzyna Socha
Department of Bromatology, Faculty of Pharmacy with the Division of Laboratory Medicine, Medical University of Białystok, Mickiewicza 2D Street, 15-222 Białystok, Poland
*
Author to whom correspondence should be addressed.
These authors contributed equally to this work.
Nutrients 2020, 12(11), 3458; https://doi.org/10.3390/nu12113458
Submission received: 3 November 2020 / Revised: 8 November 2020 / Accepted: 9 November 2020 / Published: 11 November 2020
(This article belongs to the Special Issue Nutrition in Cognitive Impairment, Dementia, and Alzheimer's Disease)

Abstract

:
Alzheimer’s disease (AD) is the most common form of dementia, and the aging of the population means that the number of cases is successively increasing. The cause of the disease has not been established, but it is suggested that many factors affect it, including nutritional aspects. As part of the work, the PubMed database has been searched, beginning from 2005, for terms related to key nutritional aspects. A diet rich in antioxidant vitamins can improve the cognitive functions of patients. Thanks to an adequate intake of B vitamins, homocysteine levels are reduced, which indirectly protects against the development of the disease. A properly balanced diet, as well as the use of appropriate supplementation, can contribute to improving the clinical condition of patients with AD.

1. Introduction

Alzheimer’s disease (AD) is one of the most common forms of dementia in the world, accounting for 2/3 of all dementias. In aging societies, the number of AD patients is predicted to increase. For example, in 2005 in Poland there were 5.9 million people over 60, and according to estimates, by 2030 this number will have increased to over 9 million [1,2]. In 2050, the disease will have affected about 115 million people worldwide [3].
The cause of AD is still unknown, but there are several hypotheses that attempt to explain the etiology of the disease. Usually, the cause is multifactorial. The gold standard in diagnosis is the accumulation of β-amyloid plaques inside the cerebral cortex and the presence of tangles of intra-nerve filamentous neurons containing tau proteins, as well as amyloid angiopathy [4]. The role of microRNA (miRNA) in neuronal cell death that is induced by β amyloid is emphasized [5]. About 95% of the cases begin at the age of 60–65 [4].
In AD, memory disorders and progressive cognitive dysfunctions occur, such as speech disorders, impairment of visual-spatial skills, and behavioral disorders [6,7]. Many symptoms, such as sleep disturbances and deterioration of mood and anxiety are revealed before clinical diagnosis of dementia. Symptoms such as confusion, aggressive behavior, and delusions may remain undiagnosed for a long time [8]. In AD, apart from the classic amnestic form, other forms are distinguished, with predominant language disorders (referred to as the speech or language variant), the visual variant, the variant with behavior disorders (frontal variant), and the form with apraxia (apractic variant) (Table 1) [9].
Currently, AD therapy uses drugs from the group of acetylcholinesterase inhibitors, such as rivastigmine and donepezil, slowing down the breakdown of acetylcholine, as well as memantine, an N-methyl-D-aspartate receptor (NMDA receptor) antagonist, whose task is to block glutamate receptors. Memantine is recommended for moderate and severe patients. Unfortunately, none of the current pharmacological treatments can stop the disease, but only decelerate its progression [10].
There are reports of common pathophysiology of two amyloidoses, AD and type 2 diabetes. In both conditions, insulin resistance, carbohydrate disorders, systemic oxidative stress, inflammation, amyloid accumulation, and cognitive decline occur. These features may indicate a common basis for these diseases. AD with concomitant type 2 diabetes is associated with so-called brain insulin resistance, which may increase the additional pathophysiological mechanisms of AD. The onset of type 2 diabetes increases the risk of AD by 50 to 100%, which is why a well-balanced diet can delay this process [11].
Optimal supply of adequate nutrients (e.g., by introducing a Mediterranean diet) or increased consumption of selected ingredients (for example, a ketogenic diet) [12], as well as avoiding toxic elements, are essential for the proper functioning of the brain and, consequently, the prevention of neurodegenerative diseases. Therefore, this publication reviews the latest literature on supporting AD therapy by taking vitamins, mainly with antioxidant properties.

2. Materials and Methods

This publication reviewed research on the importance of nutrition in AD, published from 2005 to 2020. Only human models were included in the review, with a minimum of 100 participants. In the case of several basic research papers and textbooks, reference was made to the older studies. The PubMed publication database searched for “Alzheimer’s disease”, “cognitive impairment”, “diet”, “vitamins”, “intake”, “level in serum”. The full texts of selected papers were analyzed in order to assess whether they met the adopted criteria. The publications available in a language other than English were excluded. This review presents the most important publications describing the state of the art regarding vitamin intake in AD.

3. Results and Discussion

3.1. B Vitamins

Vitamins B are a diverse group in terms of structure and function. Folates, which provide methyl groups necessary for DNA methylation, play an important role in the pathogenesis of neurological diseases [13]. The beneficial effect of vitamin B on cognitive function is associated with the metabolism of homocysteine, which contributes to increased cardiovascular risk and increased cognitive impairment [14]. In patients with AD, there is a decrease in DNA methylation. Nutrition can affect the regulation of gene expression and, thus, indirectly, can have an impact on the course of the disease. These vitamins are substances closely related to the methionine cycle [7]. Vitamin B12 is involved in the transformation of homocysteine to methionine, and vitamin B6 and folic acid are cofactors for this reaction. The deficiency of the above vitamins causes accumulation of homocysteine in the body [15]. This reduces the concentration of S-adenosylmethionine, which is a methyl donor. DNA demethylation is induced, which increases the overexpression of genes associated with the occurrence of AD [7].
Table 2 contains data on the intervention and measurement of B vitamins in AD patients.
The effect of B group vitamins (folic acid, pyridoxine and cobalamin) on homocysteine levels was confirmed in 266 participants over 70 years of age with mild cognitive impairment (MCI). Half of the participants (n = 133) were in the placebo group. The following supplementation was used: 0.8 mg folic acid, 20 mg vitamin B6, and 0.5 mg vitamin B12 for 2 years. As a result of therapy with group B vitamins, homocysteine concentration in the supplemented group was 30% lower compared to the placebo group. It was noted that patients undergoing supplementation (initial methionine concentration above 11.3 μmol/L) achieved better results in the Mini-Mental State Examination (MMSE) test, episodic memory, and semantic memory [16].
In a randomized study conducted in newly diagnosed AD patients over 60 years of age (n = 121), the effect of supplementation with folic acid on the inflammatory process and on cognition was assessed. Patients used donepezil, and for half a year they were additionally supplemented with folic acid at a dose of 1.25 mg/day. Supplemented patients obtained statistically higher results in the MMSE test compared to the control group, which did not receive folic acid. The authors emphasize that inflammation may play an important role in the interaction between AD and folic acid [17].
Different effects were noticed in other studies, where supplementation with vitamin B12 and folic acid did not reduce the development of cognitive disorders in elderly people with elevated levels of homocysteine; homocysteine concentration after 24 months of supplementation (9.3 ± 2.4 μmol/L), folate (48.0 ± 12.6 nmol/L), and active vitamin B12 (123.6 ± 13.6 pmol/L) [18].
A multicenter, double-blind study involving 2919 patients over 65 years of age, studied the effects of ingestion of 400 μg of folic acid and 500 μg of vitamin B12. The placebo and study group received an additional 15 μg of vitamin D3. The authors did not show a beneficial effect on cognitive skills. However, they emphasized that the above supplementation might slow down cognitive impairment [19].
In the study conducted by Oulhaj et al. [20], 266 participants with mild cognitive impairment were divided into a supplement group (folic acid, vitamin B6 and B12) for 2 years and a placebo group. It was shown that treatment with group B vitamins did not slow down the cognitive dysfunctions when the concentration of omega-3 fatty acids was low, whereas when the concentration was above 579 μmol/L, these vitamins inhibited the development of dementia processes, which was confirmed by Hopkins Verbal Learning Test-Revised (HVLT-DR) longitudinal scores of episodic memory. Treatment with vitamin B caused a decrease in total homocysteine levels, which resulted in a reduction in the rate of gray matter atrophy.
A cross-sectional study conducted by Meng et al. [21] on 4605 participants demonstrated that higher serum homocysteine levels were associated with an increased risk of AD. High levels of folic acid and vitamin B12 were factors protecting against cognitive impairment [21].
Ulusu et al. [22] attempted to use the determination of vitamin B12 and folic acid in a group of 290 patients as a marker of dementia. This study needs to be reconducted on a larger number of patients because a classification error was made: 5 patients had normal serum folate levels and the reference range was 4.6 to 18.7 ng/mL.
A cohort study by An et al., published in 2019 [23] involving 2533 participants with dementia or normal cognitive function, found that insufficient vitamin B12 intake contributed to cognitive decline, and adequate levels of folic acid, vitamin B6, and vitamin B12 had a beneficial effect on the cognitive resources of participants without cognitive impairment.
In summary, the effect of nutritional intervention in the form of dietary supplementation with vitamins B seems to be inconclusive, as shown by the analysis of individual publications. While it has been indicated that the measurement of serum folate and vitamin B12 levels is not a reliable test, it is suggested that high serum levels of vitamin B12 may be a protective factor.
Recommendations for the consumption of vitamin B for adults have been set at the level of the estimated average requirement (EAR) for B1, 1.1 and 1.3 mg of thiamine (for men and women, respectively); for B2, 1.1 and 0.9 mg of riboflavin; for B3, 12 and 11 mg niacin equivalent; for B6, 1.4 and 1.3 mg of vitamin B6, 320 µg of folate equivalent and 2 µg of cobalamin for both gender [24].

3.2. Antioxidant Vitamins

Increased oxidative stress is one of the factors contributing to the onset and advancement of neurodegenerative processes. Impaired antioxidant processes, excessive free radical accumulation, and age-related oxidative stress of brain tissue and lymphocytes are noted [25]. Pleiotropic mechanism of antioxidants, manifested, e.g., in a reduction of oxidized lipid membranes, reduction of nucleic acid damage or prevention of carbonylation of proteins, prevents the identification of a damaged specific pathway that is responsible for the formation of AD [26].
In a 2014 meta-analysis, levels of vitamins, microelements, and fatty acids in the blood were assessed in patients with AD and in a group of elderly people without cognitive impairment. Statistically significant vitamin deficiencies were found in A (9 tests), C (8 tests), E (20 tests), folates (31 tests), and vitamin B12 (32 tests) in the group of patients with AD, which indicates the need to provide the correct amounts of the above ingredients in diet or, in the absence of such a possibility, through supplementation. It was emphasized that AD patients have impaired systemic availability of the studied nutrients [27].

3.2.1. Vitamin A

A characteristic feature of inflammation of the nervous system is microglial activation. Inflammation occurs in chronic and acute neuropsychiatric diseases. Literature data indicates that microglia activation is known to be one of the causes of AD. Moreover, malfunctioning of the microglia may result in changes in local concentrations of retinoic acid [28]. In addition, vitamin A and its derivatives play an important role in the differentiation of nerve cells, as well as in the expression of neurotransmitters in the brain and gene expression through interaction with retinoic acid and retinoid X receptors [29].
A literature review showed that vitamin A was the least studied of the antioxidant vitamins. Vitamin A intake or dietary intake may improve cognitive function in AD patients, as shown in Table 3.
Low dietary intake of vitamin A is associated with an increased risk of dementia, which was confirmed in the group of 333 participants [30].
A study by Yuan et al. (2020) showed that higher consumption of β-carotene reduced the risk of cognitive impairment [31]. There are also studies that do not confirm the beneficial effects of this ingredient.
In a prospective study conducted by Pèneau et al., it was shown that the consumption of fruits, which are a source of β-carotene, did not reduce the risk of cognitive dysfunction [32]. Other studies on the effects of vitamin A on AD do not meet the inclusion criteria for this article.
Moreover, vitamin A is involved in neuronal differentiation and also influences the secretion of neurotransmitters in the brain [29,33].
In conclusion, it has been shown that AD patients are deficient in vitamin A, but in the case of increased consumption of β-carotene, the assessment of its effect on cognitive abilities has failed to provide definitive evidence of benefit. For example, the vitamin A intake recommendations for the population in Poland, at the level of the EAR is 630 μg of retinol equivalent/day for men, for women: 500 μg/day [24].

3.2.2. Vitamin C and E

Vitamin E can be stored in the central nervous system, where it reduces lipid peroxidation and β-amyloid deposition. Therapy with antioxidant substances should be implemented in the earliest possible stage of the disease [34].
In patients with AD, there is a correlation between blood vitamin C and vitamin E levels and dementia [35]. Ascorbic acid has a neuroprotective effect because it has the ability to scavenge free radicals, reduces β-amyloid activity, and is also involved in the chelation of iron, zinc, and copper. It is a key antioxidant of the central nervous system [36]. In a rat study, a pro-oxidative diet has been shown to increase the level of amyloid precursor protein [37].
Table 4 summarizes the data on the effect of vitamin C and E intake or supplementation on the cognitive abilities of AD patients.
In the study conducted on 613 patients, the effects of α-tocopherol (at a dose of 2000 IU/day), memantine (20 mg/day), and combinations of these substances were assessed. The observation period was 2.27 years. Vitamin E alone proved to slow down cognitive decline in patients with mild to moderate AD. Interestingly, memantine or memantine therapy, together with α-tocopherol gave less beneficial effects [38].
The results of studies on vitamin E supplementation in patients with AD are inconclusive. In a cohort study, 7540 patients over 60 years of age were treated with long-term supplementation with vitamin E (400 IU/day), selenium (200 μg/day), and vitamin E with selenium (duration of use: 5.4 ± 1.2 years), but the therapy did not prevent dementia [39]. Similar observations were obtained in a study assessing vitamin E supplementation at a dose of 300 mg and 400 mg of vitamin C on the cognitive function of the elderly with MCI (n = 256, age: 60–75 years). The intervention lasted a year and resulted in a reduction of the oxidative stress of the body, including a decrease in the level of malonic aldehyde, but did not improve cognitive abilities [40].
Higher brain γ-tocopherol levels are associated with presynaptic protein levels in the midfrontal cortex. These results suggest that vitamin E plays a significant role in maintaining presynaptic protein levels (synaptophysin, synaptotagmin, septin 5, syntaxin, synaptosomal associated protein 25 (SNAP-25), vesicle-associated membrane protein (VAMP), complexin-I, complexin-II) [41].
Patients with mild cognitive impairment (MCI, n = 166) and people diagnosed with AD (n = 168) have been shown to have significantly lower total tocopherols (6.80 μmol/mmol cholesterol and 6.40 μmol/mmol, respectively), total tocotrienols (97.28 nM/mmol cholesterol and 91.33 nM/mmol cholesterol, respectively) and total vitamin E in the blood (6.9 μmol/mmol cholesterol and 6.49 μmol/mmol), compared to persons with normal cognitive abilities (CN, n = 187), whose total tocopherols were 7.67 μmol/mmol cholesterol, total tocotrienols—118.02 nM/mmol cholesterol, total vitamin E—7.8 μmol/mmol cholesterol [42].
In a 13-year study on a group of 2533 French people aged 45 to 60, the relationship between fruit and vegetable consumption (400 g) and cognitive ability was assessed using neuropsychological tests. It was proved that the consumption of fruit and other products rich in vitamin C and vitamin E positively correlated with the results of verbal memory test, in contrast to the consumption of vegetables and vegetables rich in β-carotene [32].
Another prospective, longitudinal cohort study (n = 9250) also confirmed the beneficial effect of fruit consumption on reducing the risk of AD [43].
A review of the literature from 2012 assessing the effect of ascorbic acid intake on cognitive functions revealed ambiguous information about the effectiveness of the use of vitamin C in AD. The hypothesis was presented that the brain draws ascorbic acid from the peripheral pool to endure oxidative stress [44].
In conclusion, research results indicate that decreased vitamin E concentration may be associated with an increased risk of AD, supplementation at higher doses (2000 IU) than lower doses (400 IU) is more effective, and increased consumption of antioxidant vitamins (E and C) may help to improve cognitive abilities.
Vitamin C intake by adults should be 75 mg/day for men and 60 mg/day for women. Standards are set at the EAR level. The vitamin E intake recommendations at the level of the adequate intake (AI) is 10 μg of retinol equivalent/day for men, for women, 8 μg/day [24].

3.2.3. Vitamin D

Vitamin D also plays an important role in neurodegenerative processes. Its deficiency is a genetic risk factor for AD, Parkinson’s disease, multiple sclerosis, and vascular dementia [45].
Table 5 summarizes the research on the effects of vitamin D on the cognitive abilities of patients with AD.
In randomized, double-blinded placebo-controlled study on 2044 participants, supplementation of 1000 mg of calcium carbonate with 400 IU of vitamin D(3) did not improve cognitive impairment [46].
The effect of vitamin D levels on cognitive functions was studied in a group of 146 patients (78 women and 68 men). Among other things, brain imaging by magnetic resonance imaging was performed. It was shown that patients who obtained lower results in the MMSE test were characterized by lower serum vitamin D concentration [47].
Low blood levels of vitamin D may increase cognitive decline in people with dementia [48,49]. The InCHIANTI study, conducted on 858 patients over 65 years of age, showed that low serum 25-hydroxyvitamin levels were associated with an increased risk of cognitive impairment. When assessing the vitamin’s concentration in patients with a deficiency (below 25 nmol/L) and patients with normal levels (above 75 nmol/L), the relative risk of corrected cognitive decline was 1.6 [49].
In two cohort studies, no protective effect of high vitamin D levels on cognitive performance was observed [50,51].
Leeuw et al. [52] evaluated 28 nutritional biomarkers in blood and 5 in cerebrospinal fluid. It was shown that higher levels of w1.25 (OH) in serum were associated with impaired cognitive functions. Among other factors, the authors also mentioned low levels of S-adenosylmethionine and high levels of HDL, theobromine, cholesterol, and iron.
Vitamin D has the ability to reduce inflammation in the hippocampus and mitigate the accumulation of β-amyloid in the process of increased phagocytosis. This vitamin stabilizes the activity of calcium channels, regulating calcium homeostasis, disturbed by the deposition of β-amyloid. In addition, it increases the expression of vitamin D receptor, which further exerts an antioxidant effect [53]. It has been shown that vitamin D therapy in the course of AD prevents the synthesis of β-amyloid and promotes the clearance of this peptide from the brain [45]. A dose of 50,000 IU, used 3 times a week for 4 weeks, was safe and effectively normalized 25(OH) D levels in the blood serum of patients [54]. In a Annweiler et al. protocol (2011) [55], it was assumed that supplementation with 100,000 IU of cholecalciferol with memantine for one month enhanced cognition and memory capacity in patients with moderate AD. There is evidence that high vitamin D levels improve cognition, particularly in mild AD [56].
In addition, it is suspected that the use of vitamin D together with memantine in patients with moderate forms of the disease improves cognitive performance and memory ability by reducing the loss of neurons. The neuroprotective effects of vitamin D and memantine may be potentiated [53].
Summing up, a low level of vitamin D in the serum is associated with the risk of cognitive disorders, but supplementation with a dose of 400 IU does not bring any results (elderly people are recommended several times higher doses).
For example, the vitamin D intake recommendations for the population in Poland, at the level of the AI is 15 μg of cholecalciferol/day for men and for women [24].

3.2.4. Other Factors

Recently published systematic reviews and meta-analyses show the beneficial effects of various diets, including the Mediterranean diet and the “healthy diet”, which are based, among other things, on consuming more foods that provide the vitamins discussed in this study.
There is no correlation between meat consumption and the occurrence of cognitive impairment [57].
The Mediterranean diet is characterized by high consumption of vegetables and fruits, which are exceptionally rich in water-soluble vitamins. It has been shown that the Mediterranean diet protects against the development of cognitive impairment [58], whereas a ketogenic diet may have an impact on memory functions due to its influence on neurotransmission. In a systemic review by Grammatikopoulou et al., the application of ketogenic therapy (Medium Chain Triglicerydes, MCT supplementation), both long-term and interventional, improved cognitive functions [59].
The effect of “healthy diets” on cognition has also been studied. It has been reported that healthy eating patterns reduce the risk of dementia, while a high-fat diet and diets with a high glycemic index increase neuronal damage [60,61]. Proinflammatory diet also may increase the progression of dementia [62].
It has been proven that the use of nutritional formulas, fatty acids, ginseng, inositol, probiotics, and products rich in flavonoids may delay the progression of dementia [63,64]. On the other hand, according to Heider et al.’s (2020) systematic review and meta-analyses, supplementation of vitamin B, E, omega-3, and nutritional formulas does not improve cognitive function [65].
Table 6 summarizes the meta-analyzes in which the impact of consumption of B vitamins, vitamins C, D, and E on the health of AD patients was assessed.
In summary, the impact of most of the vitamins on cognitive function remains ambiguous. As far as B vitamins are concerned, the reasons authors give for the inefficiency of supplementation include: the insufficient size of the group to detect differences in cognitive abilities, the load on blood vessels, and the use of other drugs (which could reduce patients’ sensitivity to supplementation), no imaging of brain lesions, which is the most sensitive and reliable method of assessing changes in the brain [18], the interaction between, among others, vitamin B12 and vitamin D, which could weaken the effect of vitamin B12 [19]; in the case of the absence of expected effects of increased intake or supplementation with vitamin A (which is a fat-soluble vitamin): increased fatty acid content in patients, possibly body fat [30]; in the case of no improvement in cognitive function after an intervention involving supplementation with vitamin E, the authors point to inappropriate group selection (e.g., only one sex), too short duration of supplementation, too low doses of vitamin [39]. As for the reasons for the lack of effectiveness of vitamin D supplementation, the authors emphasize, inter alia, taking only 1 measurement, different levels of education, and groups disparate in terms of age [48].
Future research on people with Alzheimer’s disease should focus on further evaluation of the impact of vitamin intake on cognitive performance in patients in order to be able to clearly define their effect and formulate international guidelines for diet and supplementation in patients with AD. In addition, they should explain the mechanisms that are not fully understood to be able to develop appropriate supplement formulations that would be easily distributed in the patient’s body.
To promote healthy aging, governments in many countries are introducing public health support programs. For example, the Universities of the Third Age are established that support the memory and cognitive abilities of the elderly. Older people, for economic reasons, do not use all the medications recommended for their specific conditions, therefore, programs are being introduced under which seniors receive some drugs for free (in Poland, it is the “Drugs75+” program). In addition, pharmaceutical care (including care for the elderly) is developing in Europe, the aim of which is to help senior citizens in proper pharmacotherapy: assessing whether the preparations and dietary supplements used do not interact, and whether the elderly person correctly uses, e.g., an inhaler or a glucometer (people with AD may have difficulty with this). Containers are sold in which the AD caregiver can distribute medications for the entire week. In addition, insufficient intake of vitamins may be related to the general malnutrition of the elderly. There are programs in Europe to prevent malnutrition in older people, this problem affects 13.5 to 29.7% of the elderly, including PROMISS (Prevention of Malnutrition In Senior Subjects in the EU) [74].

4. Conclusions

Research suggests that a diet rich in antioxidant vitamins may slow the progression of Alzheimer’s disease. Data on supplementation with vitamins A, B, C, D, and E vitamins are ambiguous. There are no international guidelines for dietary recommendations in AD. The conducted meta-analyses suggest that vitamins A, B, C, D, and E deficiencies may increase the risk of AD, therefore, a diet rich in vitamins A, D, C, B6, B12, and folic acid, as well as supplementation with higher doses (e.g., 2000 IE of vitamin E) seems to be recommended in the prevention of Alzheimer’s disease. Adequate quantities of fruit and vegetables will supplement the deficiency of antioxidant vitamins, which are often deficient in patients with dementia. Further research should focus on resolving unclear issues.

Author Contributions

Conceptualization, A.M., A.P.-J., R.M.-Ż., and K.S.; methodology, A.M. and A.P.-J.; writing—original draft preparation, A.M. and A.P.-J.; writing—review and editing, K.S. and R.M.-Ż.; visualization, A.M. and A.P.-J.; supervision, K.S. and R.M.-Ż.; project administration, K.S.; funding acquisition, K.S. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Medical University of Białystok, grant number SUB/2/DN/20/003/2216.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Kalaria, R.N.; Maestre, G.E.; Arizaga, R.; Friedland, R.P.; Galasko, D.; Hall, K.; Luchsinger, J.A.; Ogunniyi, A.; Perry, E.K.; Potocnik, F.; et al. Alzheimer’s disease and vascular dementia in developing countries: Prevalence, management, and risk factors. Lancet Neurol. 2008, 7, 812–826. [Google Scholar] [CrossRef] [Green Version]
  2. Gaweł, M.; Potulska-Chomik, A. Choroby neurodegeneracyjne: Choroba Alzheimera i Parkinsona [in Polish, Neurodegenerative diseases: Alzheimer’s and Parkinson’s disease]. Post. Nauk Med. 2015, 28, 468–475. [Google Scholar]
  3. Alzheimer’s Association. 2014 Alzheimer’s disease facts and figures. Alzheimer Dement. 2014, 10, e47–e92. [Google Scholar] [CrossRef] [Green Version]
  4. Bird, T.D. Alzheimer Disease Overview. 2018. Available online: https://www.ncbi.nlm.nih.gov/books/NBK1161/ (accessed on 14 February 2020).
  5. Kim, Y.-J.; Kim, S.H.; Park, Y.; Park, J.; Lee, J.H.; Kim, B.C.; Song, W.K. miR-16-5p is upregulated by amyloid β deposition in Alzheimer’s disease models and induces neuronal cell apoptosis through direct targeting and suppression of BCL-2. Exp. Gerontol. 2020, 136, 110954. [Google Scholar] [CrossRef]
  6. Cummings, J.L.; Isaacson, R.S.; Schmitt, F.A.; Velting, D.M. A practical algorithm for managing Alzheimer’s disease: What, when, and why? Ann. Clin. Transl. Neurol. 2015, 2, 307–323. [Google Scholar] [CrossRef]
  7. Athanasopoulos, D.; Karagiannis, G.; Tsolaki, M. Recent Findings in Alzheimer Disease and Nutrition Focusing on Epigenetics. Adv. Nutr. 2016, 7, 917–927. [Google Scholar] [CrossRef]
  8. Atri, A. The Alzheimer’s Disease Clinical Spectrum. Med. Clin. N. Am. 2019, 103, 263–293. [Google Scholar] [CrossRef]
  9. Martyniuk-Gaca, A.; Papuć, E.; Tynecka-Turowska, M.; Rejdak, K. Choroba Alzheimera i jej rzadkie warianty oraz implikacje dla jakości życia. (In Polish: Alzheimer’s disease, its rare variants and implications for quality of life). Dobrostan a Stan Zdrowia 2016, 10, 127–138. [Google Scholar]
  10. Brunton, L.L.; Lazo, J.S.; Parker, K.L.W. Goodman and Gilman’s the Pharmacological, 1st ed.; red. Polish edition Buczko; McGraw-Hill Professional Czelej: Lublin, Poland, 2007. [Google Scholar]
  11. Talbot, K.; Wang, H.-Y.; Kazi, H.; Han, L.-Y.; Bakshi, K.P.; Stucky, A.; Fuino, R.L.; Kawaguchi, K.R.; Samoyedny, A.J.; Wilson, R.S.; et al. Demonstrated brain insulin resistance in Alzheimer’s disease patients is associated with IGF-1 resistance, IRS-1 dysregulation, and cognitive decline. J. Clin. Investig. 2012, 122, 1316–1338. [Google Scholar] [CrossRef] [Green Version]
  12. Vinciguerra, F.; Graziano, M.; Hagnäs, M.; Frittitta, L.; Tumminia, A. Influence of the Mediterranean and Ketogenic Diets on Cognitive Status and Decline: A Narrative Review. Nutrients 2020, 12, 1019. [Google Scholar] [CrossRef] [Green Version]
  13. Robinson, N.; Grabowski, P.; Rehman, I. Alzheimer’s disease pathogenesis: Is there a role for folate? Mech. Ageing Dev. 2018, 174, 86–94. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  14. Smith, P.J.; Blumenthal, J. Dietary Factors and Cognitive Decline. J. Prev. Alzheimer Dis. 2016, 3, 53–64. [Google Scholar]
  15. Mohajeri, M.H.; Troesch, B.; Weber, P. Inadequate supply of vitamins and DHA in the elderly: Implications for brain aging and Alzheimer-type dementia. Nutrients 2015, 31, 261–275. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  16. De Jager, C.A.; Oulhaj, A.; Jacoby, R.; Refsum, H.; Smith, A.D. Cognitive and clinical outcomes of homocysteine-lowering B-vitamin treatment in mild cognitive impairment: A randomized controlled trial. Int. J. Geriatr. Psychiatry 2011, 27, 592–600. [Google Scholar] [CrossRef]
  17. Chen, H.; Liu, S.; Ji, L.; Wu, T.; Ji, Y.; Zhou, Y.; Zheng, M.; Zhang, M.; Xu, W.; Huang, G. Folic Acid Supplementation Mitigates Alzheimer’s Disease by Reducing Inflammation: A Randomized Controlled Trial. Mediat. Inflamm. 2016, 2016, 1–10. [Google Scholar] [CrossRef] [Green Version]
  18. Kwok, T.; Wu, Y.; Lee, J.; Lee, R.; Yung, C.Y.; Choi, G.; Lee, V.; Harrison, J.; Lam, L.C.W.; Mok, V.C.T. A randomized placebo-controlled trial of using B vitamins to prevent cognitive decline in older mild cognitive impairment patients. Clin. Nutr. 2020, 39, 2399–2405. [Google Scholar] [CrossRef] [Green Version]
  19. Van der Zwaluw, N.L.; Dhonukshe-Rutten, R.A.M.; van Wijngaarden, J.P.; Brouwer-Brolsma, E.M.; van de Rest, O.; Veld, P.H.I.T.; Enneman, A.W.; van Dijk, S.C.; Ham, A.C.; Swart, K.M.; et al. Results of 2-year vitamin B treatment on cognitive performance: Secondary data from an RCT. Neurology 2014, 83, 2158–2166. [Google Scholar] [CrossRef]
  20. Oulhaj, A.; Jernerén, F.; Refsum, H.; Smith, A.D.; De Jager, C.A. Omega-3 Fatty Acid Status Enhances the Prevention of Cognitive Decline by B Vitamins in Mild Cognitive Impairment. J. Alzheimer Dis. 2016, 50, 547–557. [Google Scholar] [CrossRef] [Green Version]
  21. Meng, H.; Li, Y.; Zhang, W.; Zhao, Y.; Niu, X.-Y.; Guo, J. The relationship between cognitive impairment and homocysteine in a B12 and folate deficient population in China. Medicine 2019, 98, e17970. [Google Scholar] [CrossRef]
  22. Ulusu, N.N.; Yilmaz, G.; Erbayraktar, Z.; Evlice, A.T.; Aras, S.; Yener, G.; Avcı, A. A Turkish 3-center study evaluation of serum folic acid and vitamin B12 levels in Alzheimer disease. Turk. J. Med. Sci. 2015, 45, 1159–1166. [Google Scholar] [CrossRef]
  23. An, Y.; Feng, L.; Zhang, X.; Wang, Y.; Wang, Y.; Tao, L.; Qin, Z.; Xiao, R. Dietary intakes and biomarker patterns of folate, vitamin B6, and vitamin B12 can be associated with cognitive impairment by hypermethylation of redox-related genes NUDT15 and TXNRD1. Clin. Epigenetics 2019, 11, 139. [Google Scholar] [CrossRef] [Green Version]
  24. Jarosz, M. Normy Żywienia dla Populacji Polski; Instytut Żywności i Żywienia: Warsaw, Poland, 2017; (In Polish: Nutrition standards for the Polish population). [Google Scholar]
  25. Cosín-Tomàs, M.; Senserrich, J.; Arumí-Planas, M.; Alquézar, C.; Pallàs, M.; Martín-Requero, Á.; Suñol, C.; Kaliman, P.; Sanfeliu, C. Role of Resveratrol and Selenium on Oxidative Stress and Expression of Antioxidant and Anti-Aging Genes in Immortalized Lymphocytes from Alzheimer’s Disease Patients. Nutrients 2019, 11, 1764. [Google Scholar] [CrossRef] [Green Version]
  26. Eswaminathan, A.; Jicha, G.A. Nutrition and prevention of Alzheimer’s dementia. Front. Aging Neurosci. 2014, 6, 282. [Google Scholar] [CrossRef] [Green Version]
  27. Da Silva, S.L.; Vellas, B.; Elemans, S.; Luchsinger, J.; Kamphuis, P.; Yaffe, K.; Sijben, J.W.C.; Groenendijk, M.; Stijnen, T. Plasma nutrient status of patients with Alzheimer’s disease: Systematic review and meta-analysis. Alzheimer Dement. 2013, 10, 485–502. [Google Scholar] [CrossRef] [Green Version]
  28. Wołoszynowska-Fraser, M.U.; Kouchmeshky, A.; McCaffery, P. Vitamin A and Retinoic Acid in Cognition and Cognitive Disease. Annu. Rev. Nutr. 2020, 40, 247–272. [Google Scholar] [CrossRef]
  29. Sodhi, R.K.; Singh, N. Retinoids as potential targets for Alzheimer’s disease. Pharmacol. Biochem. Behav. 2014, 120, 117–123. [Google Scholar] [CrossRef]
  30. Shahar, S.; Lee, L.K.; Rajab, N.; Lim, C.L.; Harun, N.A.; Noh, M.F.N.M.; Then, S.-M.; Jamal, R. Association between vitamin A, vitamin E and apolipoprotein E status with mild cognitive impairment among elderly people in low-cost residential areas. Nutr. Neurosci. 2013, 16, 6–12. [Google Scholar] [CrossRef]
  31. Yuan, C.; Fondell, E.; Ascherio, A.; Okereke, O.I.; Grodstein, F.; Hofman, A.; Willett, W.C. Long-Term Intake of Dietary Carotenoids Is Positively Associated with Late-Life Subjective Cognitive Function in a Prospective Study in US Women. J. Nutr. 2020, 150, 1871–1879. [Google Scholar] [CrossRef]
  32. Péneau, S.; Galan, P.; Jeandel, C.; Ferry, M.; Andreeva, V.; Hercberg, S.; Kesse-Guyot, E.; The SUVI. MAX 2 Research Group. Fruit and vegetable intake and cognitive function in the SU.VI.MAX 2 prospective study. Am. J. Clin. Nutr. 2011, 94, 1295–1303. [Google Scholar] [CrossRef]
  33. Lane, M.A.; Bailey, S.J. Role of retinoid signalling in the adult brain. Prog. Neurobiol. 2005, 75, 275–293. [Google Scholar] [CrossRef]
  34. Mecocci, P.; Polidori, M.C. Antioxidant clinical trials in mild cognitive impairment and Alzheimer’s disease. Biochim. Biophys. Acta Mol. Basis Dis. 2012, 1822, 631–638. [Google Scholar] [CrossRef]
  35. Von Arnim, C.A.F.; Herbolsheimer, F.; Nikolaus, T.; Peter, R.; Biesalski, H.K.; Ludolph, A.C.; Riepe, M.; Nagel, G.; The ActiFE Ulm Study Group. Dietary Antioxidants and Dementia in a Population-Based Case-Control Study among Older People in South Germany. J. Alzheimer Dis. 2012, 31, 717–724. [Google Scholar] [CrossRef] [Green Version]
  36. Monacelli, F.; Acquarone, F.M.E.; Giannotti, C.; Borghi, R.; Nencioni, A. Vitamin C, Aging and Alzheimer’s Disease. Nutrients 2017, 9, 670. [Google Scholar] [CrossRef] [Green Version]
  37. Choudhry, F.; Howlett, D.R.; Richardson, J.C.; Francis, P.T.; Williams, R.J. Pro-oxidant diet enhances β/γ secretase-mediated APP processing in APP/PS1 transgenic mice. Neurobiol. Aging 2012, 33, 960–968. [Google Scholar] [CrossRef]
  38. Dysken, M.W.; Sano, M.; Asthana, S.; Vertrees, J.E.; Pallaki, M.; Llorente, M.; Love, S.; Schellenberg, G.D.; McCarten, J.R.; Malphurs, J.; et al. Effect of Vitamin E and Memantine on Functional Decline in Alzheimer Disease. JAMA 2014, 311, 33–44. [Google Scholar] [CrossRef]
  39. Kryscio, R.; Abner, E.L.; Caban-Holt, A.; Lovell, M.; Goodman, P.; Darke, A.K.; Yee, M.; Crowley, J.; Schmitt, F.A. Association of Antioxidant Supplement Use and Dementia in the Prevention of Alzheimer’s Disease by Vitamin E and Selenium Trial (PREADViSE). JAMA Neurol. 2017, 74, 567–573. [Google Scholar] [CrossRef]
  40. Alavi_Naeini, A.; Elmadfa, I.; Djazayery, A.; Barekatain, M.; Ghazvini, M.R.A.; Djalali, M.; Feizi, A. The effect of antioxidant vitamins E and C on cognitive performance of the elderly with mild cognitive impairment in Isfahan, Iran: A double-blind, randomized, placebo-controlled trial. Eur. J. Nutr. 2013, 53, 1255–1262. [Google Scholar] [CrossRef]
  41. De Leeuw, F.A.; Honer, W.G.; Schneider, J.A.; Morris, M.C. Brain γ-Tocopherol levels are associated with presynaptic protein levels in elderly human midfrontal cortex. J. Alzheimer Dis. 2020, 77, 619–627. [Google Scholar] [CrossRef]
  42. Mangialasche, F.; Xu, W.; Kivipelto, M.; Costanzi, E.; Ercolani, S.; Pigliautile, M.; Cecchetti, R.; Baglioni, M.; Simmons, A.; Soininen, H.; et al. Tocopherols and tocotrienols plasma levels are associated with cognitive impairment. Neurobiol. Aging 2012, 33, 2282–2290. [Google Scholar] [CrossRef]
  43. Agarwal, P.; Holland, T.M.; Wang, Y.; Holtzman, D.M.; Morris, M.C. Association of Strawberries and Anthocyanidin Intake with Alzheimer’s Dementia Risk. Nutrients 2019, 11, 3060. [Google Scholar] [CrossRef] [Green Version]
  44. Bowman, G.L. Ascorbic acid, cognitive function, and Alzheimer’s disease: A current review and future direction. BioFactors 2012, 38, 114–122. [Google Scholar] [CrossRef] [Green Version]
  45. Dursun, E.; Gezen-Ak, D. Vitamin D basis of Alzheimer’s disease: From genetics to biomarkers. Hormones 2018, 18, 7–15. [Google Scholar] [CrossRef]
  46. Rossom, R.C.; Espeland, M.A.; Manson, J.E.; Dysken, M.W.; Johnson, K.C.; Lane, R.S.; Leblanc, E.S.; Lederle, F.A.; Masaki, K.H.; Margolis, K.L. Calcium and vitamin D supplementation and cognitive impairment in the women’s health initiative. J. Am. Geriatr. Soc. 2012, 60, 2197–2205. [Google Scholar] [CrossRef]
  47. Shih, E.; Lee, W.; Hsu, J.; Wang, S.; Fuh, J.-L. Effect of vitamin D on cognitive function and white matter hyperintensity in patients with mild Alzheimer’s disease. Geriatr. Gerontol. Int. 2019, 20, 52–58. [Google Scholar] [CrossRef]
  48. Aguilar-Navarro, S.G.; Mimenza-Alvarado, A.J.; Jiménez-Castillo, G.A.; Bracho-Vela, L.A.; Yeverino-Castro, S.G.; Ávila-Funes, J.A. Association of Vitamin D with Mild Cognitive Impairment and Alzheimer’s Dementia in Older Mexican Adults. Rev. Investig. Clin. 2019, 71, 381–386. [Google Scholar] [CrossRef] [Green Version]
  49. Llewellyn, D.J. Vitamin D and Risk of Cognitive Decline in Elderly Persons. Arch. Intern. Med. 2010, 170, 1135–1141. [Google Scholar] [CrossRef]
  50. Lee, D.H.; Chon, J.; Kim, Y.; Seo, Y.K.; Park, E.J.; Won, C.W.; Soh, Y. Association between vitamin D deficiency and cognitive function in the elderly Korean population. Medicine 2020, 99, e19293. [Google Scholar] [CrossRef]
  51. Duchaine, C.S.; Talbot, D.; Nafti, M.; Giguère, Y.; Dodin, S.; Tourigny, A.; Carmichael, P.-H.; Laurin, D. Vitamin D status, cognitive decline and incident dementia: The Canadian Study of Health and Aging. Can. J. Public Health 2020, 111, 312–321. [Google Scholar] [CrossRef]
  52. De Leeuw, F.A.; Van Der Flier, W.M.; Tijms, B.M.; Scheltens, P.; Mendes, V.M.; Manadas, B.; Bierau, J.; Van Wijk, N.; Heuvel, E.G.V.D.; Mohajeri, M.H.; et al. Specific Nutritional Biomarker Profiles in Mild Cognitive Impairment and Subjective Cognitive Decline Are Associated with Clinical Progression: The NUDAD Project. J. Am. Med. Dir. Assoc. 2020, 21, 1513.e1–1513.e17. [Google Scholar] [CrossRef]
  53. Annweiler, C. Vitamin D in dementia prevention. Ann. N. Y. Acad. Sci. 2016, 1367, 57–63. [Google Scholar] [CrossRef] [Green Version]
  54. Przybelski, R.; Agrawal, S.; Krueger, D.; Engelke, J.A.; Walbrun, F.; Binkley, N. Rapid correction of low vitamin D status in nursing home residents. Osteoporos. Int. 2008, 19, 1621–1628. [Google Scholar] [CrossRef]
  55. Annweiler, C.; Fantino, B.; Parot-Schinkel, E.; Thiery, S.; Gautier, J.; Beauchet, O. Alzheimer’s disease input of vitamin D with mEmantine assay (AD-IDEA trial): Study protocol for a randomized controlled trial. Trials 2011, 12, 230. [Google Scholar] [CrossRef] [Green Version]
  56. Shen, L.; Ji, H.-F. Associations between Homocysteine, Folic Acid, Vitamin B12 and Alzheimer’s Disease: Insights from Meta-Analyses. J. Alzheimer Dis. 2015, 46, 777–790. [Google Scholar] [CrossRef]
  57. Zhang, H.; Hardie, L.J.; Bawajeeh, A.O.; Cade, J. Meat Consumption, Cognitive Function and Disorders: A Systematic Review with Narrative Synthesis and Meta-Analysis. Nutrients 2020, 12, 1528. [Google Scholar] [CrossRef]
  58. Limongi, F.; Siviero, P.; Bozanic, A.; Noale, M.; Veronese, N.; Maggi, S. The Effect of Adherence to the Mediterranean Diet on Late-Life Cognitive Disorders: A Systematic Review. J. Am. Med. Dir. Assoc. 2020, 21, 1402–1409. [Google Scholar] [CrossRef]
  59. Grammatikopoulou, M.G.; Goulis, D.G.; Gkiouras, K.; Theodoridis, X.; Gkouskou, K.K.; Evangeliou, A.; Dardiotis, E.; Bogdanos, D.P. To Keto or Not to Keto? A Systematic Review of Randomized Controlled Trials Assessing the Effects of Ketogenic Therapy on Alzheimer Disease. Adv. Nutr. 2020, 073. [Google Scholar] [CrossRef]
  60. Liu, Y.H.; Gao, X.; Na, M.; Kris-Etherton, P.; Mitchell, D.C.; Jensen, G.L. Dietary pattern, diet quality, and dementia: A systematic review and meta-analysis of prospective cohort studies. J. Alzheimer Dis. 2020, 78, 151–158. [Google Scholar] [CrossRef]
  61. Samadi, M.; Moradi, S.; Moradinazar, M.; Mostafai, R.; Pasdar, Y. Dietary pattern in relation to the risk of Alzheimer’s disease: A systematic review. Neurol. Sci. 2019, 40, 2031–2043. [Google Scholar] [CrossRef]
  62. Kheirouri, S.; Alizadeh, M. Dietary Inflammatory Potential and the Risk of Neurodegenerative Diseases in Adults. Epidemiol. Rev. 2019, 41, 109–120. [Google Scholar] [CrossRef]
  63. Moreira, S.C.; Jansen, A.K.; Silva, F.M. Dietary interventions and cognition of Alzheimer’s disease patients: A systematic review of randomized controlled trial. Dement. Neuropsychol. 2020, 14, 258–282. [Google Scholar] [CrossRef]
  64. Gkotzamanis, V.; Panagiotakos, D. Dietary interventions and cognition: A systematic review of clinical trials. Psychiatriki 2020, 31, 248–256. [Google Scholar] [CrossRef]
  65. Haider, S.; Schwarzinger, A.; Stefanac, S.; Soysal, P.; Smith, L.; Veronese, N.; Dorner, T.E.; Grabovac, I. Nutrientsal supplements for neuropsychiatric symptoms in people with dementia: A systematic review and meta-analysis. Int. J. Geriatr. Psychiatry 2020, 35, 1285–1291. [Google Scholar] [CrossRef]
  66. Zhang, C.; Luo, J.; Yuan, C.; Ding, D. Vitamin B12, B6, or folate and cognitive function in community-dwelling older adults: A systematic review and meta-analysis. J. Alzheimer Dis. 2020, 77, 791–794. [Google Scholar] [CrossRef]
  67. Clarke, R.; Bennett, D.; Parish, S.; Lewington, S.; Skeaff, M.; Eussen, S.J.P.M.; Lewerin, C.; Stott, D.J.; Armitage, J.; Hankey, G.J.; et al. Effects of homocysteine lowering with B vitamins on cognitive aging: Meta-analysis of 11 trials with cognitive data on 22,000 individuals. Am. J. Clin. Nutr. 2014, 100, 657–666. [Google Scholar] [CrossRef] [Green Version]
  68. Li, F.-J.; Shen, L.; Ji, H.-F. Dietary Intakes of Vitamin E, Vitamin C, and β-Carotene and Risk of Alzheimer’s Disease: A Meta-Analysis. J. Alzheimer Dis. 2012, 31, 253–258. [Google Scholar] [CrossRef]
  69. Chai, B.; Gao, F.; Wu, R.; Dong, T.; Gu, C.; Lin, Q.; Zhang, Y. Vitamin D deficiency as a risk factor for dementia and Alzheimer’s disease: An updated meta-analysis. BMC Neurol. 2019, 19, 284. [Google Scholar] [CrossRef]
  70. Yang, K.; Chen, J.; Li, X.; Zhou, Y. Vitamin D concentration and risk of Alzheimer disease. Medicine 2019, 98, e16804. [Google Scholar] [CrossRef]
  71. Jayedi, A.; Rashidy-Pour, A.; Shab-Bidar, S. Vitamin D status and risk of dementia and Alzheimer’s disease: A meta-analysis of dose-response. Nutr. Neurosci. 2018, 22, 750–759. [Google Scholar] [CrossRef]
  72. Ashley, S.; Bradburn, S.; Murgatroyd, C. A meta-analysis of peripheral tocopherol levels in age-related cognitive decline and Alzheimer’s disease. Nutr. Neurosci. 2019, 1–15. [Google Scholar] [CrossRef]
  73. Dong, Y.; Chen, X.; Liu, Y.; Shu, Y.; Chen, T.; Xu, L.; Li, M.; Guan, X. Do low-serum vitamin E levels increase the risk of Alzheimer disease in older people? Evidence from a meta-analysis of case-control studies. Int. J. Geriatr. Psychiatry 2017, 33, e257–e263. [Google Scholar] [CrossRef]
  74. PROMISS Nutrients for Healthy Ageing. Available online: www.promiss-vu.eu (accessed on 7 November 2020).
Table 1. Characteristics of symptoms of different variants of Alzheimer’s disease [9].
Table 1. Characteristics of symptoms of different variants of Alzheimer’s disease [9].
Language Variant/Logopenic Progressive Aphasia (LPA)Frontal VariantVisual Variant/Posterior Cortical Atrophy (PCA)Apraxic Variant
For at least the first two years, there is only an isolated language deficit of unnoticeable onset, progressive, without previous language disorders, and no other specific causes for the above disturbances.Pronounced disturbances in behavior and executive functions. These often mimic the behavioral disturbances in frontotemporal dementia.Initially, visual-spatial disorders dominate without any evident ophthalmic pathology, later psychotic symptoms appear, and in the final stage, there is a dementia syndrome with deep cognitive deficits, disturbing everyday functioning.Inability or difficulty in performing learned movements, while understanding the command, the willingness to perform it. The lack of impaired motor coordination, and weakened sensation and muscle strength.
After two years, other cognitive functions deteriorate, although the language deficit dominates and deepens much faster than the other ones. This is followed by the inability to see more than one object or one element of an object at the same time (simultaneous diagnosis), finger agnosia, page orientation disorders (right/left), and apraxia.
Table 2. Summary of studies on vitamin B intake and measurements in AD.
Table 2. Summary of studies on vitamin B intake and measurements in AD.
VitaminsType of Study Number of Participants/Age [Year] AssessmentIntervention/
Measurements
DurationResultsReferences
INTERVENTIONS
Vitamins Brandomized controlled trial266/≥70MMSE, Hopkins Verbal Learning Test, CVLT0.8 mg folic acid + 20 mg vitamin B6 + 0.5 mg vitamin B12/day2 years(+) slowing cognitive decline[16]
Folic acidrandomized controlled trial121/40–90MMSE, ADL1.25 mg of folic acid + donepezil/day6 months(+) beneficial effects in patients with AD[17]
Vitamins Brandomized placebo controlled trial279/≥65MMSE, NTB, CDR500 μg methylcobalamin + 400 μg folic acid or placebo/day2 years(-) lack of reduction of cognitive decline[18]
Vitamin B12, folic acidrandomized, placebo-controlled trial2919/74.1 ± 6.5MMSE and others tests400 μg of folic acid + 500 μg of vitamin B12/day2 years(-) lack of reduction of cognitive decline[19]
MEASUREMENTS
Folic acid, vitamins B6 and B12randomized trial266/≥70CDRfolic acid, vitamins B6 and B12 measurements in serum2 years(+) the effectiveness of vitamin B therapy when omega-3 levels are in the upper normal range[20]
B12cross-sectional study4605 (2396 with AD)/≥55China’s validation MMSEhomocysteine and B12 serum levels1 measurement(+) high folate and B12 levels are protective factors[21]
Vitamin B12 and folic acidcross-sectional study290/geriatric patientsThe authors performed testsvitamin B12 and folic acid level1 measurement(−) serum folic and vitamin B12 levels are not reliable tests for screening presymptomatic AD[22]
CONSUMPTION
Vitamins Bcohort study2533/50–70-diet intake of vitamin B12, B62.3 years(−) low intake of vitamin B12 intensify cognitive decline[23]
(+)—confirmed or (−)—no confirmed association of intervention/concentration measurement/consumption with Alzheimer’s disease, ADL, Activities of Daily Living scale; CDR, Clinical Dementia Rating scale; CVLT, California Verbal Learning Test; MMSE, Mini-Mental State Examination; NTB, Neurocognitive test Battery; TICS-M, The Modified Telephone Interview for Cognitive Status.
Table 3. Summary of studies on vitamin A intake and AD.
Table 3. Summary of studies on vitamin A intake and AD.
VitaminsStudy DesignParticipants [n]/Age [Year]AssessmentIntervention/
Measurements
DurationResultsReferences
MEASUREMENTS
Vitamin Across-sectional study333/≥60MMSE, Dementia Rating Scale, Geriatric Depression Scale, Rey Auditory Verbal Learningvitamin A serum level1 measurement(+) deficiency increases the risk of MCI[30]
CONSUMPTION
β-caroteneprospective study49,493/average: 48SCF, FFQ testlong-term intakes of carotenoids22 years(+) higher consumption of β-carotene is associated with a lower risk of cognitive performance[31]
β-carotenerandomized double-blind, placebo-controlled study2533/45–60semantic and phonemic fluency tests, executive functionsix 24-h dietary records13 years(−) consumption of fruit rich in β-carotene did not decrease the risk of developing cognitive impairment[32]
(+)—confirmed or (−)—no confirmed association of intervention/concentration measurement/consumption with Alzheimer’s disease, FFQ, Food Frequency Questionnaire; MMSE, Mini-Mental State Examination; SCF, Subjective Cognitive Function; MCI, mild cognitive impairment.
Table 4. Summary of studies on vitamin C and E consumption, supplementation, and blood measurements, and AD.
Table 4. Summary of studies on vitamin C and E consumption, supplementation, and blood measurements, and AD.
VitaminsStudy DesignParticipants [n]/Age [Year]AssessmentIntervention/MeasurementsDurationResultsReferences
INTERVENTIONS
Erandomized controlled trial613 with AD/
78.8 ± 7.1
MMSE2000 IU/day of α-tocopherol + 20 mg/day of memantine2.27 years(+) slowing functional decline[38]
Erandomized controlled trial cohort study7540/>60MIS, TICS-m, NYU Paragraph Delayed Recall400 IU of vitamin E/day and/or 200 μg of selenium/day6 years(−) supplementation does not prevent AD[39]
C and Erandomized controlled trial256/60–75MMSE300 mg of vitamin E + 400 mg of vitamin C/day1 year(−) supplementation does not prevent cognitive decline[40]
MEASUREMENTS
Ecross-sectional study113/88.5 ± 6.0MMSEα- and γ-tocopherol brain levels3 years(+) vitamin E facilitates maintenance a presynaptic proteins level[41]
Ecross-sectional study716/>65MMSEtocopherols and total vitamin E levels1 measurement(+) lower levels of total tocopherols, total tocotrienols, and total vitamin E are associated with an increased likelihood of developing AD[42]
CONSUMPTION
C and Edouble-blind, placebo-controlled2533/45–60verbal memory (RI-48 cued recall, semantic, and phonemic fluency tests) and executive function (trail-making and forward and backward digit span tests)intake of fruit and vegetable intake (24-h dietary records)13 years(+) improvement of verbal memory[32]
Cprospective, longitudinal cohort study925/58–98FFQintake of strawberries (vitamin C)6.7 years(+) possible reduction of AD[43]
(+)—confirmed or (−)—no confirmed association of intervention/concentration measurement/consumption with Alzheimer’s disease, FFQ, Food Frequency Questionnaire; MIS, Memory Impairment Screen; MMSE, Mini-Mental State Examination; NYU, New York University; TICS-m, modified Telephone Interview for Cognitive Status.
Table 5. Summary of studies on vitamin D supplementation and blood measurements in AD.
Table 5. Summary of studies on vitamin D supplementation and blood measurements in AD.
VitaminStudy DesignParticipants [n]/Age [Year]AssessmentIntervention/
Measurement
DurationResultsReference
INTERVENTION
Drandomized double-blinded placebo-controlled trial4144/>65WHIMS, MMSE1000 mg of calcium carbonate + 400 IU of
vitamin D(3)
8 years(−) no association between treatment and incident cognitive impairment[46]
MEASUREMENTS
Dcross-sectional study146/79.1 ± 7.0MMSE25(OH)D serum level1 measurement(+) higher serum 25(OH)D level are associated with MMSE score[47]
Dcross-sectional study208/79 ± 1MMSE, CDRserum 25-hidroxyvitamin D (25(OH)D) levels1 year(+) low vitamin D are associated with cognitive decline[48]
Drandomized controlled trial1154/79.7 ± 8.4MMSEserum levels of 25(OH)D8 years(+) low levels are associated with cognitive decline[49]
Dcohort study2990/76.5 ± 3.9MMSE-KC, Test Digit Span25(OH)D serum level2 years(−) no relationship between vitamin D levels and cognition[50]
Dcohort study661/≥65 3MSPlasma 25-hydroxyvitamin D10 years(+) higher 25(OH)D concentrations increase risk of dementia and AD in women[51]
Dretrospective, longitudinal cohort study299/61 ± 10, 66 ± 8MMSE28 nutritional biomarkers in blood and 5 in cerebrospinal fluid, including 1.25(OH)22–5 years(+) high vitamin D is associated with cognitive decline[52]
(+)—confirmed or (−)—no confirmed association of intervention/concentration measurement/consumption with Alzheimer’s disease, 3MS, Modified Mini-Mental State; CDR, Clinical Dementia Rating scale; MMSE, Mini-Mental State Examination; MMSE-KC, Mini-Mental State Examination in the Korean version; WHIMS, Women’s Health Initiative Memory Study.
Table 6. A review of meta-analyses and systematic reviews of recent years on the effects of nutrition on AD.
Table 6. A review of meta-analyses and systematic reviews of recent years on the effects of nutrition on AD.
FactorsType of AnalysisResultsYearReferences
VITAMINS
Vitamins B6, B12, and folatesystematic review and meta-analysis: 8 cross-sectional, 13 longitudinalNo link between the concentration of vitamin B in the blood and the risk of cognitive disorders.2020[66]
Vitamins Bmeta-analysis: 11 large studiesLowering homocysteine levels through vitamin B supplementation does not improve cognitive function.2016[67]
Folic acid, vitamin B12meta-analysis: 68 studiesHigh homocysteine levels and low levels of folic acid and vitamin B12 increase the risk of AD.2015[56]
Vitamins C, E, and β-carotenemeta-analysis: 7 studiesConsumption of vitamins E, C, and β-carotene reduces the risk of AD. Vitamin E is the strongest predictor.2012[68]
Vitamin Dmeta-analysis: 12 prospective cohort studies, 4 cross-sectional studiesVitamin D deficiency increases the risk of dementia and AD.2019[69]
Vitamin Dmeta-analysis: 6 prospective cohort studiesSerum vitamin D deficiency is associated with the risk of AD.2019[70]
Vitamin Dmeta-analysis: 7 prospective cohort studies, 1 retrospective cohort studyHigher levels of 25(OH)D reduce the risk of AD.2019[71]
Vitamin Emeta-analysis: 31 observational studies and clinical trialLower α-tocopherol levels increase the risk of AD and MCI.2019[72]
Vitamin Emeta-analysis: 17 studiesLower vitamin E levels are associated with the risk of AD.2018[73]
OTHER NUTRITIONAL FACTORS
Fatty acids, foods rich in polyphenolssystematic review: 10 randomized clinical trialsFatty acids and products rich in flavonoid are a protective factor in the development of dementia2020[64]
Ketogenic dietsystematic review: 10 randomized clinical trialsMCT and other ketogenic therapies improve cognition2020[59]
Meat consumptionsystematic review and meta-analysis (PRISMA): 10 studiesNo association between meat consumption and cognitive impairment2020[57]
Mediterranean dietsystematic review: 7 randomized clinical trials, 38 longitudinal studiesThe diet is a protective factor to cognitive decline2020[58]
Nutritional formulas, fatty acid, ginseng, inositol, probioticssystematic review: 32 randomized clinical trialsInterventions protect or delay the progression of cognitive decline2020[63]
Supplementation (vitamins B, vitamins E, omega-3 fatty acids, nutritional formulas)systematic review and meta-analyses (PRISMA): 4 randomized clinical trialsNo significant link between supplementation and cognitive function2020[65]
Traditional, healthy dietsystematic review and meta-analysis: 12 studiesHealthy diet reduces the risk of dementia2020[60]
Pro-inflammatory dietsystematic review and meta-analysis (PRISMA): 10 randomized clinical trialsInflammatory diet is associated with cognitive decline2019[62]
Traditional, healthy diet and high-fat, high-glycemic dietsystematic reviews and meta-analyses (PRISMA): 26 studiesHealthy diet protects against the development of AD, unhealthy diet increases neurodegeneration2019[61]
MCT, Medium Chain Trigliceryde.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Mielech, A.; Puścion-Jakubik, A.; Markiewicz-Żukowska, R.; Socha, K. Vitamins in Alzheimer’s Disease—Review of the Latest Reports. Nutrients 2020, 12, 3458. https://doi.org/10.3390/nu12113458

AMA Style

Mielech A, Puścion-Jakubik A, Markiewicz-Żukowska R, Socha K. Vitamins in Alzheimer’s Disease—Review of the Latest Reports. Nutrients. 2020; 12(11):3458. https://doi.org/10.3390/nu12113458

Chicago/Turabian Style

Mielech, Anita, Anna Puścion-Jakubik, Renata Markiewicz-Żukowska, and Katarzyna Socha. 2020. "Vitamins in Alzheimer’s Disease—Review of the Latest Reports" Nutrients 12, no. 11: 3458. https://doi.org/10.3390/nu12113458

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