Premutation in the Fragile X Mental Retardation 1 (FMR1) Gene Affects Maternal Zn-milk and Perinatal Brain Bioenergetics and Scaffolding

Eleonora Napoli, Catherine Ross-Inta, Gyu Song, Sarah Wong, Randi Hagerman, Louise W Gane, Jennifer T Smilowitz, Flora Tassone, Cecilia Giulivi, Eleonora Napoli, Catherine Ross-Inta, Gyu Song, Sarah Wong, Randi Hagerman, Louise W Gane, Jennifer T Smilowitz, Flora Tassone, Cecilia Giulivi

Abstract

Fragile X premutation alleles have 55-200 CGG repeats in the 5' UTR of the FMR1 gene. Altered zinc (Zn) homeostasis has been reported in fibroblasts from >60 years old premutation carriers, in which Zn supplementation significantly restored Zn-dependent mitochondrial protein import/processing and function. Given that mitochondria play a critical role in synaptic transmission, brain function, and cognition, we tested FMRP protein expression, brain bioenergetics, and expression of the Zn-dependent synaptic scaffolding protein SH3 and multiple ankyrin repeat domains 3 (Shank3) in a knock-in (KI) premutation mouse model with 180 CGG repeats. Mitochondrial outcomes correlated with FMRP protein expression (but not FMR1 gene expression) in KI mice and human fibroblasts from carriers of the pre- and full-mutation. Significant deficits in brain bioenergetics, Zn levels, and Shank3 protein expression were observed in the Zn-rich regions KI hippocampus and cerebellum at PND21, with some of these effects lasting into adulthood (PND210). A strong genotype × age interaction was observed for most of the outcomes tested in hippocampus and cerebellum, whereas in cortex, age played a major role. Given that the most significant effects were observed at the end of the lactation period, we hypothesized that KI milk might have a role at compounding the deleterious effects on the FMR1 genetic background. A higher gene expression of ZnT4 and ZnT6, Zn transporters abundant in brain and lactating mammary glands, was observed in the latter tissue of KI dams. A cross-fostering experiment allowed improving cortex bioenergetics in KI pups nursing on WT milk. Conversely, WT pups nursing on KI milk showed deficits in hippocampus and cerebellum bioenergetics. A highly significant milk type × genotype interaction was observed for all three-brain regions, being cortex the most influenced. Finally, lower milk-Zn levels were recorded in milk from lactating women carrying the premutation as well as other Zn-related outcomes (Zn-dependent alkaline phosphatase activity and lactose biosynthesis-whose limiting step is the Zn-dependent β-1,4-galactosyltransferase). In premutation carriers, altered Zn homeostasis, brain bioenergetics and Shank3 levels could be compounded by Zn-deficient milk, increasing the risk of developing emotional and neurological/cognitive problems and/or FXTAS later in life.

Keywords: FMR1; Shank3; bioenergetics; brain; milk; mitochondria; premutation; zinc.

Figures

Figure 1
Figure 1
Changes in FMRP protein expression in brain from WT and KI mice. (A) Representative Western blots of FMRP and actin protein expression levels in hippocampus and cortex of WT and KI mice. Tubulin was used as loading control. FMRP protein levels of KI mice hippocampus and cerebellum were 37 and 45% of WT at post-natal day (PND) 9, and 41 and 50% of WT at post-natal day 21. In cortex, FMRP protein levels in KI mice were 38% of WT at post-natal day 9. (B) Time-dependent changes in FMRP protein levels in hippocampus, cerebellum, and cortex respectively. Data are reported as mean ± SEM, n = 3–5 per genotype per time point. Statistical analysis was performed by Two-way ANOVA. Post-hoc analysis performed by Tukey's HSD test revealed significant differences between WT and KI, indicated in the figure by asterisk as follows. Hippocampus: ****p < 0.0001; Cerebellum: *p = 0.0110; Cortex: **p = 0.0003, *p = 0.0363, ***p = 0.0001. Statistically significant differences among time points are indicated by letters with the following p values. p = 0.0029 (a), p < 0.0001 (b), p = 0.0009 (c), p = 0.0260 (d), p = 0.0101 (e), p = 0.0033 (f), p = 0.0020 (g). For more statistical details on the genotype, age, and genotype × age effect see Table 3. (C) Correlation between FMRP and actin in the same brain areas. AUD, Arbitrary Units of Densitometry.
Figure 2
Figure 2
Brain bioenergetics of KI mice during neurodevelopment and adulthood. Mitochondria were isolated from cortex, cerebellum, and hippocampus of WT and KI pups as described in the Methods section. Activities of NADH oxidase, succinate oxidase, and cytochrome c oxidase (CCO), and respiratory control ratio (RCR) were evaluated at PND0 (cerebellum and cortex only, due to the scarcity of hippocampal tissue), PND7, PND21, and PND210. Data are reported as mean ± SEM, n = 3–7 per genotype per time point. Statistical analysis was performed by Two-way ANOVA. Post-hoc analysis performed by Tukey's HSD test revealed significant differences between WT and KI indicated by asterisks as follows NADH oxidase: *p = 0.0468; Succinate oxidase: *p = 0.0466 at PND21, *p = 0.0423 at PND210; Cytochrome c oxidase: *p = 0.0484; Coupling: **p = 0.001, ***p = 0.0004, *p = 0.0496. Letters indicate statistically significant differences amongst time points as follows. NADH oxidase: p = 0.0291 (a), p = 0.0033 (b), p = 0.0166 (c), p = 0.0320 (d), p = 0.0036 (e), p = 0.0001 (f), p = 0.0113 (g), p = 0.0032 (h), p = 0.0004 (i); Succinate oxidase: p < 0.0001 (a), p = 0.0002 (b), p = 0.0240 (c), p = 0.0434 (d), p = 0.0283 (e), p = 0.0009 (f), p = 0.0250 (g), p = 0.0013 (h); Cytochrome c oxidase: p = 0.0053 (a), p = 0.0035 (b). Coupling: p = 0.0003 (a), p = 0.0010 (b). Further statistical details on the genotype, age, and genotype × age effect can be found in the legend of Table 3.
Figure 3
Figure 3
Protein expression levels of mitochondrial ATPase β-subunit in brain from KI mice. ATPase β-subunit protein expression (normalized to GAPDH) was evaluated at PND9, 21 and 210 in hippocampus, cerebellum, and cortex lysates. Asterisk in the immunoblot image of hippocampus denotes a 90 CGG KI sample. Densitometry data for this sample have not been taken into account for averages calculation of KI. Cortex samples were ran in two different gels (PND9 and PND21-210) and are shown separately. Data are reported as mean ± SEM, n = 3–5 per genotype per time point, ran in triplicates. Statistical analysis was performed by Two-way ANOVA. Post-hoc analysis was performed by Tukey's HSD test for multiple comparisons. Significant difference between WT and KI are indicated by asterisks as follows. ****p < 0.0001; *p < 0.0492. Statistically significant differences among time points are indicated by letters as follows: p < 0.0001 (a, c), p = 0.0087 (b), p = 0.0222 (d). Further, statistical details on the genotype, age, and genotype × age effect can be found in Table 3. AUD, Arbitrary Units of Densitometry.
Figure 4
Figure 4
Correlation between mitochondrial outcomes and FMRP or FMR1 expression levels. The correlation between FMRP and FMR1 expression with a mitochondrial outcome (i.e., coupling between ATP synthesis and electron transfer or RCRu) was carried out using human primary dermal from controls, premutation carriers (105–180 CGG), and full mutation carriers (>200 CGG repeats). RCRu, FMR1, and FMRP levels were expressed as percentage of control values. Data are shown as mean ± SEM for controls, permutation, and full mutation carriers.
Figure 5
Figure 5
Protein expression of precursor and mature CCOIV in brains from KI mice. Representative Western blots of CCOIV (precursor and mature proteins) in hippocampus (A), cerebellum (B) and cortex (C) of WT and KI mice. The densitometry for all the samples is also shown. Data were expressed as Arbitrary Densitometry Units and reported as mean ± SEM. Mature form of CCO4 was normalized by GAPDH. Data are reported as mean ± SEM, n = 3–5 per genotype per time point, ran in triplicates. Statistical analysis was performed by Two-way ANOVA, followed by Tukey post-hoc test for multiple comparisons. Statistically significant differences between WT and KI are indicated by asterisks as follows. Hippocampus: *p = 0.0416; ***p = 0.0002; ****p < 0.0001. Cerebellum: *p = 0.0212 for mature CCO at PND21; *p = 0.0197 for P:M at PND21. Statistically significant differences among time points are indicated by letters as follows. Hippocampus: p = 0.0002 (a), p = 0.0172 (b), p = 0.0013 (c), p = 0.0047 (d), p = 0.0086 (e), p < 0.0001 (f), p = 0.0003 (g). Cerebellum: p = 0.0459 (a). Cortex: p = 0.0391 (a), p = 0.0113 (b). Further statistical details on the genotype, age, and genotype × age effect can be found in Table 3. (D) Uncropped version of the Western blot image showing the intensity and mobility of the precursor band relative to the mature protein. The observed molecular weights for the CCOIV precursor and mature forms were, respectively, 20.0 and 17.5 kDa, as extrapolated by the Molecular Weight markers with the use of the Carestream software. This 2.5 kDa difference was close to the theoretical calculated molecular weight (2.4 kDa) of the 22 residues of amino acids (MLATRVFSLVGKRAISTSVCVR) present in the precursor form, which is cleaved by mitochondrial matrix peptidases (Isaya et al., 1991) to produce the mature mitochondrial form of the protein (UniProtKB P13073). Asterisk in the immunoblot image of hippocampus (panels A and D) denotes a 90 CGG KI sample. Densitometry data for this sample have not been included in the averages for KI. For cortex, samples were run in two separate gels (PND9 and PND21-210). AUD, Arbitrary Units of Densitometry.
Figure 6
Figure 6
Brain Zn concentrations in WT and KI mice. Total Zn levels were measured in whole hippocampus, cerebellum and cortex homogenates from WT and KI mice. Data are shown as mean ± SEM, n = 3–6 individuals per genotype, per time point, ran individually. Statistical analysis was performed by Two-way ANOVA, followed by Tukey's HSD post-hoc test for multiple comparisons. Statistically significant differences between WT and KI at individual time-points are indicated by asterisks as follows: Hippocampus: *p = 0.0425; **p = 0.0070; Cerebellum: *p = 0.0158; ***p = 0.0003. Statistically significant differences among time points are indicated by letters as follows: p = 0.0190 (a), p = 0.0144 (b), p = 0.0003 (c). Further statistical details on the genotype, age, and genotype × age effect can be found in Table 3.
Figure 7
Figure 7
Correlations among Zn levels, selected mitochondrial outcomes, and protein expression of FMRP and Shank3, in brain of WT and KI mice. A significant direct correlation was noted between average Zn levels and activities of CCO and CS and RCR for all the time points analyzed (A). CCO and CS activities were expressed as nmol × (min × mg protein)−1. CCO activity has been normalized to citrate synthase and multiplied by 100. All the other mitochondrial outcomes did not show a correlation with Zn. A direct correlation was also observed between Zn levels and both FMRP and Shank3 protein expression (individual values) in brains of WT and KI when all time points were combined (B,C). Dotted lines illustrate the 95% CI obtained with WT and KI values.
Figure 8
Figure 8
Brain Shank3 protein levels in WT and KI mice. (A) Western blots for Shank3 protein expression levels in hippocampus, cerebellum, and cortex were performed at PND 9, 21, and 210 from WT (white) and KI (black) mice as described in details in the Materials and Methods Section. Further, images taken with acetone-precipitated samples and normalized by either tubulin or GADPH are found under Supplementary Figure 1. (B) Densitometry data is shown as mean ± SEM, n = 4–5 individuals per genotype, per time point, ran individually. Statistical analysis was performed by two-way ANOVA, followed by Tukey's HSD post-hoc test for multiple comparisons. Statistically significant differences between WT and KI are indicated by asterisks as follows: Hippocampus: **p = 0.0054; Cerebellum: ****p < 0.0001. Statistically significant differences among time points are indicated by letters as follows: p = 0.0074 (a), p = 0.0267 (b), p < 0.0001 (c, e, f), p = 0.0217(d), p = 0.0041 (g). Further, statistical details on the genotype, age, and genotype × age effect can be found in Table 3. (C) Correlation between FMRP levels and Shank3 protein expression in brain of WT and KI mice. Both proteins have been normalized by tubulin, used as loading control. Shown are individual data points collected in cerebellum, hippocampus, and cortex of WT and KI mice at PND9, PND21, and PND210. PND210 is shown in a separate plot due to the difference in protein expression of both FMRP and Shank3 at this time point, relative to the previous ones. Dotted lines are 95% CI constructed with WT and KI values. AUD, Arbitrary Units of Densitometry.
Figure 9
Figure 9
ZnT4 and ZnT6 gene expression in mammary glands from lactating WT and KI dams. Representative image of the PCR products (obtained by RT-qPCR as described under Methods) that were separated in a 1.3% agarose gel and visualized with ethidium bromide (A). The fragments exhibited the expected sizes of 585 bp (ZnT4) and 537 bp (ZnT6). GAPDH PCR product was separated in a 3% agarose gel electrophoresis with ethidium bromide resulting in a product of the expected of 107 bp (not shown). Intensities of ZnT4 and ZnT6 bands were obtained in a Kodak Imager were normalized by that of GAPDH(B). Bars represent averages ± SEM of 4 individuals/genotype. Statistical analysis was carried out with the Student's t-test between WT and KI. *p < 0.05.
Figure 10
Figure 10
Changes in mitochondrial outcomes in hippocampus, cerebellum and cortex of suckling WT and KI pups nursed on WT or KI dams. At birth, KI pups (n = 24) and WT pups (n = 24) were foster-nursed either on KI dams or WT dams, six pups on each dam. After 21 days, mitochondria were isolated from cortex, cerebellum, and hippocampus and activities of NADH oxidase, succinate oxidase, cytochrome c oxidase, and citrate synthase and RCR were evaluated as described in the Methods section. wt/wt = WT pups nursing on WT milk; ki/ki = KI pups nursing on KI milk. Circles represent WT pups, squares represent KI pups. White symbols represent WT milk and black symbols represent KI milk. Upward or downward arrows represent improvement or worsening, respectively, upon nursing on WT milk or KI milk. Activities (NADH oxidase, succinate oxidase, cytochrome c oxidase) were expressed as nmol × (min × mg protein)−1, normalized by citrate synthase activity and multiplied by 1000. Data are shown as mean ± SEM (from technical replicates of pooled samples). Statistical analysis was performed by Two-way ANOVA, followed by Tukey's post-hoc test for multiple comparisons. The p values are as follows. Hippocampus: p < 0.0001 (a, b, c, g, i, p, q), p = 0.0004 (d, o), p = 0.0002 (e, f), p = 0.0025 (h), p = 0.0008 (j), p = 0.0027 (k), p = 0.0006 (l), p = 0.0262 (m), p = 0.0082 (n); Cerebellum: p = 0.0296 (a), p = 0.0301 (b), p = 0.0039 (c), p < 0.0001 (d, g, h, i, j, k, l, m, n, o, p), p = 0.0023 (e); p = 0.0086 (f); Cortex: p < 0.0001 (a, b, c, d, e, f), p = 0.0047 (g), p = 0.0468 (h). Further statistical details on the genotype, age, and genotype × age effect can be found in Table 4.
Figure 11
Figure 11
Role of milk Zn in brain bioenergetics and synapse scaffolding in premutation carriers. Milk-Zn concentrations followed a pseudo-first order kinetics over the lactation period (A). Milk Zn concentrations from control and premutation donors decreased with the post-partum period, ranging from 30 ± 3 μM at 6–8 weeks to 15 ± 2 μM at 18–20 weeks. The LN values of the Zn concentrations (in μM) were plotted against lactation week. The published values were obtained from Nagra et al. (Nagra, 1989) and shown as white squares. Regression parameters are as follows: for published [Zn], r2 = 0.997, y = −0.09x + 4.3; for this study, r2 = 0.647, y = −0.07x + 4.0. P value for slopes = 0.458; P value for intercepts = 0.986. ALP activity [expressed as nmol × (min × mg protein)−1] in human breast milk from control donors increased linearly from 8 to 18–20 weeks (B). The average ALP activity of controls was 3.3 ± 0.3 nmol × (min × mg protein)−1 or 46 ± 1 μmol × (min × l)−1 and within reported values [40 to 60 μmol × (min × l)−1; (Worth et al., ; Coburn et al., 1992)]. Linear regression analyses were performed with control values (reported as mean ± SEM) for both outcomes (r2 = 0.638 and 0.639 for Zn concentration and ALP activity, respectively). The 95% CIs performed with control values are shown with dotted lines. Individual premutation carriers' values for Zn concentrations and ALP activities are shown in Table 5. ALP activities in controls were not affected by vitamin and mineral supplementation [supplemented: 3.7 ± 0.4 vs. not supplemented: 3.2 ± 0.4 nmol × (min × mg protein)−1; p = 0.624]. A model integrating experimental data based on altered Zn homeostasis, mitochondrial dysfunction [current study and (Ross-Inta et al., ; Napoli et al., 2011)], and the putative effect of defective synaptic scaffolding in premutation individuals is shown in (C). During lactation, Zn uptake is exerted by ZIP5, 8 and 10, then excreted to milk via ZnT4/ZnT2 with a lower flux through ZnT6 (ZnT10) in the Golgi [Adapted from (Kelleher et al., 2012)]. The presence of the CGG expansion in FMR1, as it is the case with premutation carriers, may ensue in a lower expression of FMRP, which may affect the cytoskeleton (e.g., actin, Shank3), and the expression or function of membrane transporter such as ZnT4/T6. This situation would ensue in the suboptimal delivery of Zn into milk, mitochondrial dysfunction, and synaptic dysregulation. The arrows indicate the direction of Zn transport from maternal blood to milk.

References

    1. Bañez-Coronel M., Porta S., Kagerbauer B., Mateu-Huertas E., Pantano L., Ferrer I., et al. . (2012). A pathogenic mechanism in Huntington's disease involves small CAG-repeated RNAs with neurotoxic activity. PLoS Genet. 8:e1002481. 10.1371/journal.pgen.1002481
    1. Bardoni B., Mandel J. L. (2002). Advances in understanding of fragile X pathogenesis and FMRP function, and in identification of X linked mental retardation genes. Curr. Opin. Genet. Dev. 12, 284–293. 10.1016/S0959-437X(02)00300-3
    1. Battistella G., Niederhauser J., Fornari E., Hippolyte L., Gronchi Perrin A., Lesca G., et al. . (2013). Brain structure in asymptomatic FMR1 premutation carriers at risk for fragile X-associated tremor/ataxia syndrome. Neurobiol. Aging 34, 1700–1707. 10.1016/j.neurobiolaging.2012.12.001
    1. Beach R. S., Gershwin M. E., Hurley L. S. (1982). Gestational zinc deprivation in mice: persistence of immunodeficiency for three generations. Science 218, 469–471. 10.1126/science.7123244
    1. Belanger M., Magistretti P. J. (2009). The role of astroglia in neuroprotection. Dialogues Clin. Neurosci. 11, 281–295.
    1. Boldogh I. R., Pon L. A. (2006). Interactions of mitochondria with the actin cytoskeleton. Biochim. Biophys. Acta 1763, 450–462. 10.1016/j.bbamcr.2006.02.014
    1. Brand M. D. (1990). The contribution of the leak of protons across the mitochondrial inner membrane to standard metabolic rate. J. Theor. Biol. 145, 267–286. 10.1016/S0022-5193(05)80131-6
    1. Brody T. (1999). Nutritional Biochemistry. San Diego, CA: Academic Press.
    1. Brown V., Jin P., Ceman S., Darnell J. C., O'Donnell W. T., Tenenbaum S. A., et al. . (2001). Microarray identification of FMRP-associated brain mRNAs and altered mRNA translational profiles in fragile X syndrome. Cell 107, 477–487. 10.1016/S0092-8674(01)00568-2
    1. Butterworth R. F., Heroux M. (1989). Effect of pyrithiamine treatment and subsequent thiamine rehabilitation on regional cerebral amino acids and thiamine-dependent enzymes. J. Neurochem. 52, 1079–1084. 10.1111/j.1471-4159.1989.tb01850.x
    1. Castets M., Schaeffer C., Bechara E., Schenck A., Khandjian E. W., Luche S., et al. . (2005). FMRP interferes with the Rac1 pathway and controls actin cytoskeleton dynamics in murine fibroblasts. Hum. Mol. Genet. 14, 835–844. 10.1093/hmg/ddi077
    1. Chanda R., Owen E. C., Cramond B. (1951). The composition of human milk with special reference to the relation between phosphorus partition and phosphatase and to the partition of certain vitamins. Br. J. Nutr. 5, 228–242. 10.1079/BJN19510029
    1. Chonchaiya W., Tardif T., Mai X., Xu L., Li M., Kaciroti N., et al. . (2013). Developmental trends in auditory processing can provide early predictions of language acquisition in young infants. Dev. Sci. 16, 159–172. 10.1111/desc.12012
    1. Coburn S. P., Mahuren J. D., Pauly T. A., Ericson K. L., Townsend D. W. (1992). Alkaline phosphatase activity and pyridoxal phosphate concentrations in the milk of various species. J. Nutr. 122, 2348–2353.
    1. Darnell J. C., Richter J. D. (2012). Cytoplasmic RNA-binding proteins and the control of complex brain function. Cold Spring Harb. Perspect. Biol. 4:a012344. 10.1101/cshperspect.a012344
    1. Darnell J. C., Van Driesche S. J., Zhang C., Hung K. Y., Mele A., Fraser C. E., et al. . (2011). FMRP stalls ribosomal translocation on mRNAs linked to synaptic function and autism. Cell 146, 247–261. 10.1016/j.cell.2011.06.013
    1. D'Hulst C., Heulens I., Brouwer J. R., Willemsen R., De Geest N., Reeve S. P., et al. . (2009). Expression of the GABAergic system in animal models for fragile X syndrome and fragile X associated tremor/ataxia syndrome (FXTAS). Brain Res. 1253, 176–183. 10.1016/j.brainres.2008.11.075
    1. Dugina V., Zwaenepoel I., Gabbiani G., Clement S., Chaponnier C. (2009). Beta and gamma-cytoplasmic actins display distinct distribution and functional diversity. J. Cell Sci. 122, 2980–2988. 10.1242/jcs.041970
    1. Elfering S. L., Haynes V. L., Traaseth N. J., Ettl A., Giulivi C. (2004). Aspects, mechanism, and biological relevance of mitochondrial protein nitration sustained by mitochondrial nitric oxide synthase. Am. J. Physiol. Heart Circ. Physiol. 286, H22–H29. 10.1152/ajpheart.00766.2003
    1. Farzin F., Perry H., Hessl D., Loesch D., Cohen J., Bacalman S., et al. . (2006). Autism spectrum disorders and attention-deficit/hyperactivity disorder in boys with the fragile X premutation. J. Dev. Behav. Pediatr. 27, S137–S144. 10.1097/00004703-200604002-00012
    1. Fransson G. B., Lonnerdal B. (1984). Iron, copper, zinc, calcium, and magnesium in human milk fat. Am. J. Clin. Nutr. 39, 185–189.
    1. Fujisawa Y., Napoli E., Wong S., Song G., Yamaguchi R., Matsui T., et al. . (2015). Impact of a novel homozygous mutation in nicotinamide nucleotide transhydrogenase on mitochondrial DNA integrity in a case of familial glucocorticoid deficiency. Proc. Natl. Acad. Sci. U.S.A. 3, 70–78. 10.1016/j.bbacli.2014.12.003
    1. Garcia-Nonell C., Ratera E. R., Harris S., Hessl D., Ono M. Y., Tartaglia N., et al. . (2008). Secondary medical diagnosis in fragile X syndrome with and without autism spectrum disorder. Am. J. Med. Genet. A 146A, 1911–1916. 10.1002/ajmg.a.32290
    1. Grabrucker S., Jannetti L., Eckert M., Gaub S., Chhabra R., Pfaender S., et al. . (2014). Zinc deficiency dysregulates the synaptic ProSAP/Shank scaffold and might contribute to autism spectrum disorders. Brain 137, 137–152. 10.1093/brain/awt303
    1. Greco C. M., Berman R. F., Martin R. M., Tassone F., Schwartz P. H., Chang A., et al. . (2006). Neuropathology of fragile X-associated tremor/ataxia syndrome (FXTAS). Brain 129, 243–255. 10.1093/brain/awh683
    1. Hagerman R., Hagerman P. (2013). Advances in clinical and molecular understanding of the FMR1 premutation and fragile X-associated tremor/ataxia syndrome. Lancet Neurol. 12, 786–798. 10.1016/S1474-4422(13)70125-X
    1. Hagerman R., Hoem G., Hagerman P. (2010). Fragile X and autism: intertwined at the molecular level leading to targeted treatments. Mol. Autism 1:12. 10.1186/2040-2392-1-12
    1. Han K., Holder J. L., Jr., Schaaf C. P., Lu H., Chen H., Kang H., et al. . (2013). SHANK3 overexpression causes manic-like behaviour with unique pharmacogenetic properties. Nature 503, 72–77. 10.1038/nature12630
    1. Hara Y., Yuk F., Puri R., Janssen W. G., Rapp P. R., Morrison J. H. (2014). Presynaptic mitochondrial morphology in monkey prefrontal cortex correlates with working memory and is improved with estrogen treatment. Proc. Natl. Acad. Sci. U.S.A. 111, 486–491. 10.1073/pnas.1311310110
    1. Harris S. W., Hessl D., Goodlin-Jones B., Ferranti J., Bacalman S., Barbato I., et al. . (2008). Autism profiles of males with fragile X syndrome. Am. J. Ment. Retard. 113, 427–438. 10.1352/2008.113:427-438
    1. Hatch A. L., Gurel P. S., Higgs H. N. (2014). Novel roles for actin in mitochondrial fission. J. Cell Sci. 127, 4549–4560. 10.1242/jcs.153791
    1. Haynes V., Traaseth N. J., Elfering S., Fujisawa Y., Giulivi C. (2010). Nitration of specific tyrosines in FoF1 ATP synthase and activity loss in aging. Am. J. Physiol. Endocrinol. Metab. 298, E978–E987. 10.1152/ajpendo.00739.2009
    1. Higuchi R., Vevea J. D., Swayne T. C., Chojnowski R., Hill V., Boldogh I. R., et al. . (2013). Actin dynamics affect mitochondrial quality control and aging in budding yeast. Curr. Biol. 23, 2417–2422. 10.1016/j.cub.2013.10.022
    1. Ho E., Courtemanche C., Ames B. N. (2003). Zinc deficiency induces oxidative DNA damage and increases p53 expression in human lung fibroblasts. J. Nutr. 133, 2543–2548.
    1. Huang L., Gitschier J. (1997). A novel gene involved in zinc transport is deficient in the lethal milk mouse. Nat. Genet. 17, 292–297. 10.1038/ng1197-292
    1. Hunsaker M. R., Arque G., Berman R. F., Willemsen R., Hukema R. K. (2012). Mouse models of the fragile x premutation and the fragile X associated tremor/ataxia syndrome. Results Probl. Cell Differ. 54, 255–269. 10.1007/978-3-642-21649-7_14
    1. Hunsaker M. R., von Leden R. E., Ta B. T., Goodrich-Hunsaker N. J., Arque G., Kim K., et al. . (2011). Motor deficits on a ladder rung task in male and female adolescent and adult CGG knock-in mice. Behav. Brain Res. 222, 117–121. 10.1016/j.bbr.2011.03.039
    1. Hunsaker M. R., Wenzel H. J., Willemsen R., Berman R. F. (2009). Progressive spatial processing deficits in a mouse model of the fragile X premutation. Behav. Neurosci. 123, 1315–1324. 10.1037/a0017616
    1. Isaya G., Kalousek F., Fenton W. A., Rosenberg L. E. (1991). Cleavage of precursors by the mitochondrial processing peptidase requires a compatible mature protein or an intermediate octapeptide. J. Cell Biol. 113, 65–76. 10.1083/jcb.113.1.65
    1. Iwata-Ichikawa E., Kondo Y., Miyazaki I., Asanuma M., Ogawa N. (1999). Glial cells protect neurons against oxidative stress via transcriptional up-regulation of the glutathione synthesis. J. Neurochem. 72, 2334–2344. 10.1046/j.1471-4159.1999.0722334.x
    1. Jagoda C. A., Rillema J. A. (1991). Temporal effect of prolactin on the activities of lactose synthetase, alpha-lactalbumin, and galactosyl transferase in mouse mammary gland explants. Proc. Soc. Exp. Biol. Med. 197, 431–434. 10.3181/00379727-197-43278
    1. Jekabsons M. B., Nicholls D. G. (2004). In situ respiration and bioenergetic status of mitochondria in primary cerebellar granule neuronal cultures exposed continuously to glutamate. J. Biol. Chem. 279, 32989–33000. 10.1074/jbc.M401540200
    1. Kaplan E. S., Cao Z., Hulsizer S., Tassone F., Berman R. F., Hagerman P. J., et al. . (2012). Early mitochondrial abnormalities in hippocampal neurons cultured from Fmr1 pre-mutation mouse model. J. Neurochem. 123, 613–621. 10.1111/j.1471-4159.2012.07936.x
    1. Kelleher S. L., Velasquez V., Croxford T. P., McCormick N. H., Lopez V., MacDavid J. (2012). Mapping the zinc-transporting system in mammary cells: molecular analysis reveals a phenotype-dependent zinc-transporting network during lactation. J. Cell. Physiol. 227, 1761–1770. 10.1002/jcp.22900
    1. Kenneson A., Zhang F., Hagedorn C. H., Warren S. T. (2001). Reduced FMRP and increased FMR1 transcription is proportionally associated with CGG repeat number in intermediate-length and premutation carriers. Hum. Mol. Genet. 10, 1449–1454. 10.1093/hmg/10.14.1449
    1. Kim S. Y., Hashimoto R., Tassone F., Simon T. J., Rivera S. M. (2013). Altered neural activity of magnitude estimation processing in adults with the fragile X premutation. J. Psychiatr. Res. 47, 1909–1916. 10.1016/j.jpsychires.2013.08.014
    1. Knapka J. J., Smith K. P., Judge F. J. (1974). Effect of open and closed formula rations on the performance of three strains of laboratory mice. Lab. Anim. Sci. 24, 480–487.
    1. Kogan C. S., Turk J., Hagerman R. J., Cornish K. M. (2008). Impact of the Fragile X mental retardation 1 (FMR1) gene premutation on neuropsychiatric functioning in adult males without fragile X-associated Tremor/Ataxia syndrome: a controlled study. Am. J. Med. Genet. B Neuropsychiatr. Genet. 147B, 859–872. 10.1002/ajmg.b.30685
    1. Korobova F., Ramabhadran V., Higgs H. N. (2013). An actin-dependent step in mitochondrial fission mediated by the ER-associated formin INF2. Science 339, 464–467. 10.1126/science.1228360
    1. Kouser M., Speed H. E., Dewey C. M., Reimers J. M., Widman A. J., Gupta N., et al. . (2013). Loss of predominant Shank3 isoforms results in hippocampus-dependent impairments in behavior and synaptic transmission. J. Neurosci. 33, 18448–18468. 10.1523/JNEUROSCI.3017-13.2013
    1. Krebs N. F., Hambidge K. M. (2007). Complementary feeding: clinically relevant factors affecting timing and composition. Am. J. Clin. Nutr. 85, 639S–645S.
    1. Krebs N. F., Reidinger C. J., Hartley S., Robertson A. D., Hambidge K. M. (1995). Zinc supplementation during lactation: effects on maternal status and milk zinc concentrations. Am. J. Clin. Nutr. 61, 1030–1036.
    1. Kusano H., Shimizu S., Koya R. C., Fujita H., Kamada S., Kuzumaki N., et al. . (2000). Human gelsolin prevents apoptosis by inhibiting apoptotic mitochondrial changes via closing VDAC. Oncogene 19, 4807–4814. 10.1038/sj.onc.1203868
    1. Li J., Li Q., Xie C., Zhou H., Wang Y., Zhang N., et al. . (2004). Beta-actin is required for mitochondria clustering and ROS generation in TNF-induced, caspase-independent cell death. J. Cell Sci. 117, 4673–4680. 10.1242/jcs.01339
    1. Li S., Xu S., Roelofs B. A., Boyman L., Lederer W. J., Sesaki H., et al. . (2015). Transient assembly of F-actin on the outer mitochondrial membrane contributes to mitochondrial fission. J. Cell Biol. 208, 109–123. 10.1083/jcb.201404050
    1. Ludwig A. L., Espinal G. M., Pretto D. I., Jamal A. L., Arque G., Tassone F., et al. . (2014). CNS expression of murine fragile X protein (FMRP) as a function of CGG-repeat size. Hum. Mol. Genet. 23, 3228–3238. 10.1093/hmg/ddu032
    1. Luecke R. W., Fraker P. J. (1979). The effect of varying dietary zinc levels on growth and antibody-mediated response in two strains of mice. J. Nutr. 109, 1373–1376.
    1. McCormick N. H., Kelleher S. L. (2012). ZnT4 provides zinc to zinc-dependent proteins in the trans-Golgi network critical for cell function and Zn export in mammary epithelial cells. Am. J. Physiol. Cell Physiol. 303, C291–C297. 10.1152/ajpcell.00443.2011
    1. Menalled L., El-Khodor B. F., Patry M., Suarez-Farinas M., Orenstein S. J., Zahasky B., et al. . (2009). Systematic behavioral evaluation of Huntington's disease transgenic and knock-in mouse models. Neurobiol. Dis. 35, 319–336. 10.1016/j.nbd.2009.05.007
    1. Mohammad M. A., Hadsell D. L., Haymond M. W. (2012). Gene regulation of UDP-galactose synthesis and transport: potential rate-limiting processes in initiation of milk production in humans. Am. J. Physiol. Endocrinol. Metab. 303, E365–E376. 10.1152/ajpendo.00175.2012
    1. Moore M. E., Moran J. R., Greene H. L. (1984). Zinc supplementation in lactating women: evidence for mammary control of zinc secretion. J. Pediatr. 105, 600–602. 10.1016/S0022-3476(84)80430-8
    1. Nagra S. A. (1989). Longitudinal study in biochemical composition of human milk during first year of lactation. J. Trop. Pediatr. 35, 126–128. 10.1093/tropej/35.3.126
    1. Napoli E., Hung C., Wong S., Giulivi C. (2013). Toxicity of the flame-retardant BDE-49 on brain mitochondria and neuronal progenitor striatal cells enhanced by a PTEN-deficient background. Toxicol. Sci. 132, 196–210. 10.1093/toxsci/kfs339
    1. Napoli E., Ross-Inta C., Wong S., Hung C., Fujisawa Y., Sakaguchi D., et al. . (2012). Mitochondrial dysfunction in Pten haplo-insufficient mice with social deficits and repetitive behavior: interplay between Pten and p53. PLoS ONE 7:e42504. 10.1371/journal.pone.0042504
    1. Napoli E., Ross-Inta C., Wong S., Omanska-Klusek A., Barrow C., Iwahashi C., et al. . (2011). Altered zinc transport disrupts mitochondrial protein processing/import in fragile X-associated tremor/ataxia syndrome. Hum. Mol. Genet. 20, 3079–3092. 10.1093/hmg/ddr211
    1. Navarro D., Zwingmann C., Butterworth R. F. (2008). Region-selective alterations of glucose oxidation and amino acid synthesis in the thiamine-deficient rat brain: a re-evaluation using 1H/13C nuclear magnetic resonance spectroscopy. J. Neurochem. 106, 603–612. 10.1111/j.1471-4159.2008.05410.x
    1. Nguyen L. S., Lepleux M., Makhlouf M., Martin C., Fregeac J., Siquier-Pernet K., et al. . (2016). Profiling olfactory stem cells from living patients identifies miRNAs relevant for autism pathophysiology. Mol. Autism 7, 1. 10.1186/s13229-015-0064-6
    1. Nicholls D. G. (2008). Oxidative stress and energy crises in neuronal dysfunction. Ann. N.Y. Acad. Sci. 1147, 53–60. 10.1196/annals.1427.002
    1. Nolze A., Schneider J., Keil R., Lederer M., Huttelmaier S., Kessels M. M., et al. . (2013). FMRP regulates actin filament organization via the armadillo protein p0071. RNA 19, 1483–1496. 10.1261/rna.037945.112
    1. Oh S. Y., He F., Krans A., Frazer M., Taylor J. P., Paulson H. L., et al. . (2015). RAN translation at CGG repeats induces ubiquitin proteasome system impairment in models of fragile X-associated tremor ataxia syndrome. Hum. Mol. Genet. 24, 4317–4326. 10.1093/hmg/ddv165
    1. Picard M., McEwen B. S. (2014). Mitochondria impact brain function and cognition. Proc. Natl. Acad. Sci. U.S.A. 111, 7–8. 10.1073/pnas.1321881111
    1. Pieretti M., Zhang F. P., Fu Y. H., Warren S. T., Oostra B. A., Caskey C. T., et al. . (1991). Absence of expression of the FMR-1 gene in fragile X syndrome. Cell 66, 817–822. 10.1016/0092-8674(91)90125-I
    1. Rendon A., Masmoudi A. (1985). Purification of non-synaptic and synaptic mitochondria and plasma membranes from rat brain by a rapid Percoll gradient procedure. J. Neurosci. Methods 14, 41–51. 10.1016/0165-0270(85)90113-X
    1. Ross-Inta C., Omanska-Klusek A., Wong S., Barrow C., Garcia-Arocena D., Iwahashi C., et al. . (2010). Evidence of mitochondrial dysfunction in fragile X-associated tremor/ataxia syndrome. Biochem. J. 429, 545–552. 10.1042/BJ20091960
    1. Russo A. J., Devito R. (2011). Analysis of copper and zinc plasma concentration and the efficacy of zinc therapy in individuals with asperger's syndrome, pervasive developmental disorder not otherwise specified (PDD-NOS) and autism. Biomark. Insights 6, 127–133. 10.4137/BMI.S7286
    1. Sawashita J., Takeda A., Okada S. (1997). Change of zinc distribution in rat brain with increasing age. Brain Res. Dev. Brain Res. 102, 295–298. 10.1016/S0165-3806(97)00107-7
    1. Selkoe D. J. (2002). Alzheimer's disease is a synaptic failure. Science 298, 789–791. 10.1126/science.1074069
    1. Sellier C., Freyermuth F., Tabet R., Tran T., He F., Ruffenach F., et al. . (2013). Sequestration of DROSHA and DGCR8 by expanded CGG RNA repeats alters microRNA processing in fragile X-associated tremor/ataxia syndrome. Cell Rep. 3, 869–880. 10.1016/j.celrep.2013.02.004
    1. Sellier C., Rau F., Liu Y., Tassone F., Hukema R. K., Gattoni R., et al. . (2010). Sam68 sequestration and partial loss of function are associated with splicing alterations in FXTAS patients. EMBO J. 29, 1248–1261. 10.1038/emboj.2010.21
    1. Stewart R. A., Platou E., Kelly V. J. (1958). The alkaline phosphatase content of human milk. J. Biol. Chem. 232, 777–784.
    1. Sutcliffe J. S., Nelson D. L., Zhang F., Pieretti M., Caskey C. T., Saxe D., et al. . (1992). DNA methylation represses FMR-1 transcription in fragile X syndrome. Hum. Mol. Genet. 1, 397–400. 10.1093/hmg/1.6.397
    1. Tassone F., Beilina A., Carosi C., Albertosi S., Bagni C., Li L., et al. . (2007). Elevated FMR1 mRNA in premutation carriers is due to increased transcription. RNA 13, 555–562. 10.1261/rna.280807
    1. Tassone F., Greco C. M., Hunsaker M. R., Seritan A. L., Berman R. F., Gane L. W., et al. . (2012). Neuropathological, clinical and molecular pathology in female fragile X premutation carriers with and without FXTAS. Genes Brain Behav. 11, 577–585. 10.1111/j.1601-183X.2012.00779.x
    1. Tassone F., Hagerman P. J. (2003). Expression of the FMR1 gene. Cytogenet. Genome Res. 100, 124–128. 10.1159/000072846
    1. Tassone F., Hagerman R. J., Loesch D. Z., Lachiewicz A., Taylor A. K., Hagerman P. J. (2000a). Fragile X males with unmethylated, full mutation trinucleotide repeat expansions have elevated levels of FMR1 messenger RNA. Am. J. Med. Genet. 94, 232–236. 10.1002/1096-8628(20000918)94:3<232::AID-AJMG9>;2-H
    1. Tassone F., Hagerman R. J., Taylor A. K., Gane L. W., Godfrey T. E., Hagerman P. J. (2000b). Elevated levels of FMR1 mRNA in carrier males: a new mechanism of involvement in the fragile-X syndrome. Am. J. Hum. Genet. 66, 6–15. 10.1086/302720
    1. Tassone F., Hagerman R. J., Taylor A. K., Mills J. B., Harris S. W., Gane L. W., et al. . (2000c). Clinical involvement and protein expression in individuals with the FMR1 premutation. Am. J. Med. Genet. 91, 144–152. 10.1002/(SICI)1096-8628(20000313)91:2<144::AID-AJMG14>;2-V
    1. Tassone F., Iwahashi C., Hagerman P. J. (2004). FMR1 RNA within the intranuclear inclusions of fragile X-associated tremor/ataxia syndrome (FXTAS). RNA Biol. 1, 103–105. 10.4161/rna.1.2.1035
    1. Todd P. K., Oh S. Y., Krans A., He F., Sellier C., Frazer M., et al. . (2013). CGG repeat-associated translation mediates neurodegeneration in fragile X tremor ataxia syndrome. Neuron 78, 440–455. 10.1016/j.neuron.2013.03.026
    1. Tokatlidis K., Vial S., Luciano P., Vergnolle M., Clémence S. (2000). Membrane protein import in yeast mitochondria. Biochem. Soc. Trans. 28, 495–499. 10.1042/bst0280495
    1. Tudehope D. I. (2013). Human milk and the nutritional needs of preterm infants. J. Pediatr. 162, S17–S25. 10.1016/j.jpeds.2012.11.049
    1. Umeta M., West C. E., Haidar J., Deurenberg P., Hautvast J. G. (2000). Zinc supplementation and stunted infants in Ethiopia: a randomised controlled trial. Lancet 355, 2021–2026. 10.1016/S0140-6736(00)02348-5
    1. Van Dam D., Errijgers V., Kooy R. F., Willemsen R., Mientjes E., Oostra B. A., et al. . (2005). Cognitive decline, neuromotor and behavioural disturbances in a mouse model for fragile-X-associated tremor/ataxia syndrome (FXTAS). Behav. Brain Res. 162, 233–239. 10.1016/j.bbr.2005.03.007
    1. Wang X., Xu Q., Bey A. L., Lee Y., Jiang Y. H. (2014). Transcriptional and functional complexity of Shank3 provides a molecular framework to understand the phenotypic heterogeneity of SHANK3 causing autism and Shank3 mutant mice. Mol. Autism 5, 30. 10.1186/2040-2392-5-30
    1. Weiler I. J., Irwin S. A., Klintsova A. Y., Spencer C. M., Brazelton A. D., Miyashiro K., et al. . (1997). Fragile X mental retardation protein is translated near synapses in response to neurotransmitter activation. Proc. Natl. Acad. Sci. U.S.A. 94, 5395–5400. 10.1073/pnas.94.10.5395
    1. Wenzel H. J., Hunsaker M. R., Greco C. M., Willemsen R., Berman R. F. (2010). Ubiquitin-positive intranuclear inclusions in neuronal and glial cells in a mouse model of the fragile X premutation. Brain Res. 1318, 155–166. 10.1016/j.brainres.2009.12.077
    1. Winarni T. I., Chonchaiya W., Sumekar T. A., Ashwood P., Morales G. M., Tassone F., et al. . (2012). Immune-mediated disorders among women carriers of fragile X premutation alleles. Am. J. Med. Genet. A 158A, 2473–2481. 10.1002/ajmg.a.35569
    1. Wong L. M., Goodrich-Hunsaker N. J., McLennan Y., Tassone F., Harvey D., Rivera S. M., et al. . (2012). Young adult male carriers of the fragile X premutation exhibit genetically modulated impairments in visuospatial tasks controlled for psychomotor speed. J. Neurodev. Disord. 4:26. 10.1186/1866-1955-4-26
    1. Worth G. K., Retallack R. W., Gutteridge D. H., Jefferies M., Kent J., Smith M. (1981). Serum and milk alkaline phosphatase in human lactation. Clin. Chim. Acta 115, 171–177. 10.1016/0009-8981(81)90073-5
    1. Xu X., Forbes J. G., Colombini M. (2001). Actin modulates the gating of Neurospora crassa VDAC. J. Membr. Biol. 180, 73–81. 10.1007/s002320010060
    1. Yadava N., Nicholls D. G. (2007). Spare respiratory capacity rather than oxidative stress regulates glutamate excitotoxicity after partial respiratory inhibition of mitochondrial complex I with rotenone. J. Neurosci. 27, 7310–7317. 10.1523/JNEUROSCI.0212-07.2007
    1. Yasuda H., Yoshida K., Yasuda Y., Tsutsui T. (2011). Infantile zinc deficiency: association with autism spectrum disorders. Sci. Rep. 1:129. 10.1038/srep00129
    1. Zalewski P., Truong-Tran A., Lincoln S., Ward D., Shankar A., Coyle P., et al. . (2006). Use of a zinc fluorophore to measure labile pools of zinc in body fluids and cell-conditioned media. BioTechniques 40, 509–520. 10.2144/06404RR02
    1. Zalfa F., Giorgi M., Primerano B., Moro A., Di Penta A., Reis S., et al. . (2003). The fragile X syndrome protein FMRP associates with BC1 RNA and regulates the translation of specific mRNAs at synapses. Cell 112, 317–327. 10.1016/S0092-8674(03)00079-5
    1. Zingerevich C., Greiss-Hess L., Lemons-Chitwood K., Harris S. W., Hessl D., Cook K., et al. . (2009). Motor abilities of children diagnosed with fragile X syndrome with and without autism. J. Intellect. Disabil. Res. 53, 11–18. 10.1111/j.1365-2788.2008.01107.x
    1. Zongaro S., Hukema R., D'Antoni S., Davidovic L., Barbry P., Catania M. V., et al. . (2013). The 3′ UTR of FMR1 mRNA is a target of miR-101, miR-129-5p and miR-221: implications for the molecular pathology of FXTAS at the synapse. Hum. Mol. Genet. 22, 1971–1982. 10.1093/hmg/ddt044

Source: PubMed

3
订阅