The neuropeptide AgRP is essential for maintaining systemic energy homeostasis. In the current study, we show that hypothalamic Foxi2, as a novel regulator of nutrient sensing, controls systemic energy metabolism by specifically stimulating AgRP expression. Foxi2 was highly expressed in the hypothalamus, and its expression was induced by fasting. Immunofluorescence assays demonstrated that Foxi2 was localized in AgRP neurons. We stereotactically injected adeno-associated virus to selectively overexpress Foxi2 in AgRP-IRES-Cre mice and found that Foxi2 overexpression in AgRP neurons specifically increased AgRP expression, thereby increasing food intake and reducing energy expenditure, subsequently leading to obesity and insulin resistance. Mechanistically, Foxi2 stimulated AgRP expression by directly binding to it and activating its transcription. Furthermore, Foxi2 overexpression activated AgRP neuron activity, as revealed by whole-cell patch-clamp experiments. Conversely, global Foxi2-mutant mice became leaner with age and were resistant to high-fat diet–induced obesity and metabolic disturbances. Collectively, our data suggest that Foxi2 plays an important role in controlling energy metabolism by regulating AgRP expression.

Obesity is a major health problem that results from a disequilibrium between energy intake and energy expenditure (1,2). Energy metabolism is tightly controlled by the hypothalamus. The arcuate nucleus (ARC), located on the floor of the third ventricle, is the chief hypothalamic area involved in the control of food intake, and it contains two interconnected groups of neurons that release agouti-related protein (AgRP) and neuropeptide Y (NPY), which stimulate food intake, and the anorexigenic peptides proopiomelanocortin (POMC) and cocaine- and amphetamine-regulated transcript (CART) (3,4). AgRP neurons have emerged as key mediators of appetite behavior. Pharmacogenetic or optogenetic activation of AgRP neurons in mice results in rapid stimulation of food intake, even in caloric-replete states (5). These effects of AgRP neurons on food intake are mediated by the release of AgRP, NPY, and γ-aminobutyric acid (6,7). POMC is the precursor of α-melanocyte–stimulating hormone, and the anorexigenic effect of melanocortin is mediated by two receptors, MC3R and MC4R, on second-order neurons (8). AgRP influences food intake mainly through the competitive antagonism of central melanocortin receptors.

In addition to food intake, AgRP neurons regulate energy expenditure and substrate utilization. Acute activation of AgRP neurons reduces energy expenditure, while mice lacking AgRP neurons have enhanced energy expenditure (9,10), indicating a direct relationship between AgRP neurons and brown fat function.

The forkhead class of transcription factors is characterized by a 110–amino acid monomeric DNA-binding domain (11). The forkhead box i (Foxi) subfamily comprises three members (Foxi1, Foxi2, and Foxi3) (12). Mice with a Foxi1-null mutation exhibit defects in inner ear formation, resulting in hearing and balance impairment (13). Foxi3-mutant mice are not viable and display severe branchial arch–derived facial skeleton defects (14). Foxi2 is mainly expressed in the developing forebrain and has a significant role in embryonic development, especially brain development (15). However, whether and how Foxi2 regulates systemic metabolism remains unknown.

In the current study, we show that Foxi2 exerts its effects on food intake and energy expenditure through induction of AgRP expression. Our study reveals a novel regulator of nutrient sensing and energy homeostasis.

Animal Treatment

Global Foxi2-mutant mice were obtained from the Cyagen Biosciences (stock no. 270004). AgRP-IRES-Cre mice were purchased from The Jackson Laboratory (stock no. 012899). Wild-type (WT) C57BL/6N mice were purchased from SPF (Beijing) Biotechnology Co., Ltd. All animals were housed in 12-h light and 12-h dark photoperiods. For high-fat diet (HFD) experiments, 6–8-week-old male mice were fed an HFD with 60% kcal from fat (D12492; Research Diets, New Brunswick, NJ) for 3–4 months. All animal experiments were reviewed and approved by the institutional animal care and use committee of Tianjin Medical University.

Glucose and Insulin Tolerance Tests

Glucose intolerance tests (GTTs) and insulin tolerance tests (ITTs) were performed as described previously (16,17). For GTT, mice were fasted overnight and injected intraperitoneally with d-glucose (2 g/kg). For ITT, mice were fasted for 6 h and injected intraperitoneally with human insulin (0.75 units/kg; Solarbio). Blood glucose concentrations were measured from the tail vein at indicated times.

Food Intake Studies

For the daily food intake assay, mice were individually housed in cages for 5 days. Food intake and spillage were measured during the last 4 days (18). For fasting-induced refeeding assays, the single-housed mice were deprived of food overnight, then a preweighed amount of food was placed in cages, and food intake was calculated for 24 h (19).

Metabolic Parameter Measurements

Body weight was measured weekly. Body temperature was measured by rectal thermometer. Body composition (fat and lean mass) was determined by MRI (EchoMRI-Combo-700). Indirect calorimetry was performed according to standard methods using metabolic chambers (TSE Systems, Columbus Instruments). Energy expenditure and locomotor activity were monitored for 48 h in metabolic cages. Before data collection, the mice were acclimated to their individual housing and calorimetry chamber for 24 h. Room temperature was ∼23°C, and light/dark cycles were 12 h (18,20). The rate of energy expenditure was calculated using a modified Weir equation (energy expenditure [kcal/h] = 60 × [0.003941 × VO2 + 0.001106 × VCO2]) and respiratory quotient (RQ) as VCO2/VO2.

Preparation of Expression Plasmids, Recombinant Adenoviruses, and Lentiviruses

The full-length mouse Foxi2 gene containing a Flag-tag at the C-terminus was cloned into pcDNA3.1. The Flag-tagged Foxi2 gene was constructed into pAdM-FH-GFP, and recombinant adenovirus expressing Foxi2 was generated by Vigene Biosciences (Jinan, China). For Foxi2 knockdown, shRNA sequences were synthesized by Tsingke Biotechnology Co., Ltd. (Beijing, China) and constructed into lentivirus plasmids pLKO.1-TRC cloning vector (no. 10878; Addgene); the primer sequences are shown in Supplementary Table 1. Lentiviruses were generated as described previously (21). The concentrated viruses were used to infect GT1-7 cells.

Chromatin Immunoprecipitation Assay

Briefly, GT1-7 cells were infected with Ad-GFP or Ad-Foxi2 for 48 h and then were harvested for chromatin immunoprecipitation (ChIP) assay as described previously (22). Primer sequences for quantitative PCR are listed in Supplementary Table 1.

Luciferase Assays

pGL3-AgRP promoter (−1795 to 255 base pairs [bp]) was constructed with NheI/XhoI sites. A series of 5′ deletion constructs of AgRP (−1795 luciferase [Luc], −545Luc, −345Luc, and −195Luc) were prepared by PCR using −1795Luc as a template. Primer sequences are shown in Supplementary Table 1. Luciferase was measured as described previously (23).

Stereotactic Viral Injection of Adeno-Associated Virus Into the ARC

The pAAV-EF1a-DIO-EGFP-WPRE (adeno-associated virus [AAV]-GFP) and pAAV-EF1a-DIO-Foxi2-P2A-EGFP-WPRE (AAV-Foxi2) viruses (Obio Technology Co., Ltd., Shanghai, China) were injected bilaterally into the ARC of AgRP-IRES-Cre male mice as described previously (19). Briefly, using a stereotactic device, AAVs were injected slowly with a glass micropipette at a rate of 30–40 nL/min (∼2 × 1013 vg/mL) for 10 min according to the following coordinates: −1.50 mm posterior to the bregma, ±0.25 mm lateral to the midline, and −5.80 mm below the surface of the skull. The injector (P1 Technologies) was left in position for an additional 10 min to permit diffusion. After surgery, mice were individually housed and allowed 2 weeks to recover before the start of studies. Each group contained seven mice. The anatomical accuracy and effectiveness of stereotactic viral injection were verified under fluorescence microscopy (for electrophysiology-related studies) or by immunofluorescence assay and quantitative PCR (for anatomy and in vivo studies). The “missed” mice were excluded from data analyses.

Histology Analysis

For hematoxylin-eosin (H-E) staining, immunohistochemistry, and oil red O staining, inguinal white adipose tissue (iWAT), epididymal white adipose tissue (eWAT), brown adipose tissue (BAT), and liver tissue were treated as described previously (24). Immunofluorescence staining was performed using Foxi2 (sc-515114; Santa Cruz Biotechnology) and anti-AgRP (ab254558; Abcam) as described previously (19).

Western Blot Analysis

Immunoblotting was performed using the following primary antibodies: Foxi2 (sc-515114; Santa Cruz Biotechnology), β-tubulin (A12289; ABclonal), β-actin (AC026; ABclonal), p-AKT (Ser473) (#9271; Cell Signaling Technology [CST]), AKT (#9272; CST), UCP1 (#14670; CST); ATGL (A5126; ABclonal), p-HSL (Ser563) (#4139; CST), HSL (WL02643; Wanleibio), and tyrosine hydroxylase (WL01820; Wanleibio).

RNA Extraction and Quantitative PCR Analysis

Total RNA was extracted from tissues or cells with TRIzol reagent (Invitrogen). Real-time PCR was performed as described previously (25). The related primers are shown in Supplementary Table 1.

Metabolite Measurement

Serum triacylglycerol (TG) and total cholesterol (TC) levels were determined using an automated Monarch device (Tianjin Medical University General Hospital). Hepatic concentrations of TG and TC were quantified with enzymatic assays (Applygen Technologies, Beijing, China) according to the manufacturer’s instructions. Serum insulin levels were determined using an ELISA kit (#EZRMI-13K; EMD Millipore).

Electrophysiological Slice Recordings

Mice fed a standard chow diet were fasted 24 h, then sacrificed for electrophysiological recordings as described previously (26). Briefly, mice were anesthetized with isoflurane, and brains were removed and submerged in ice-cold artificial cerebrospinal fluid (ACSF) containing 120 mmol/L NaCl, 2.5 mmol/L KCl, 2 mmol/L CaCl2 ⋅ 2H2O, 1.25 mmol/L KH2PO4, 2 mmol/L MgSO4 ⋅ 7H2O, 26 mmol/L NaHCO3, and 10 mmol/L d-glucose saturated in 95% O2/5% CO2. Coronal sections containing the ARC (300 μm) were cut using a Leica VT1200S vibratome and bathed in oxygenated ACSF (33°C) for 30 min, then slices were incubated in oxygenated ACSF at room temperature for at least 1 h before recording. Slices were removed to a recording chamber and bathed in oxygenated ACSF at a flow rate of 1.5–2 mL/min at room temperature. The pipette solution for whole-cell recording contained 140 mmol/L potassium gluconate, 2 mmol/L MgCl2, 10 mmol/L HEPES, 10 mmol/L KCl, 5 mmol/L EGTA, 1 mmol/L CaCl2, and 2 mmol/L Mg ATP (pH 7.3). The osmolarity of the internal solution was adjusted to ∼290 mmol/kg. Recording electrodes had resistances of 3–5 mol/LΩ when filled with potassium gluconate internal solution. We determined the location of AAV-infected AgRP neurons in ARC visualized by fluorescence microscope (40× infrared lens) (BX51WI; Olympus) with blue excitation block, at which time the light source was switched to bright field to record neuronal discharge. For current-clamp experiments, the resting membrane potential was measured immediately after achieving whole-cell configuration, and at least 5 min was allowed for equilibration between pipette solution and cell to permit membrane conductance to stabilize. Current injection started at −10 pA and was gradually increased to 40 pA to assess neuronal excitability and membrane properties. The current injection step was held at 10 pA, and the action potential (AP) recording time was 500 ms. For spontaneous excitatory postsynaptic current (sEPSC) experiments, cells were held in voltage-clamp mode at −70 mV for 5 min in the presence of 100 μmol/L picrotoxin to eliminate ionotropic GABAergic transmission. Electrophysiological signals were recorded using an Axopatch 700B amplifier (Molecular Devices), low-pass filtered at 2–5 kHz, and analyzed offline on a personal computer running the pCLAMP program (Molecular Devices). The series resistance in whole-cell patch-clamp recording was <30 mol/LΩ and remained stable during the course of the experiments.

Statistics

All data are presented as mean ± SEM of three independent experiments. Results were analyzed using Microsoft Excel and/or GraphPad Prism software. Statistical significance was determined by unpaired two-tailed Student t test and one-way ANOVA followed by Tukey multiple comparison test. For the curves of body weight and tolerance tests, a t test was used to compare differences between groups of mice at each time point. For energy expenditure, an ANCOVA was performed using SPSS software. ANCOVA with fat and lean mass as well as body weight as covariates was performed for energy expenditure to statistically adjust for body composition differences between groups. Adjusted marginal means and partial η2 values as approximations of effect size were calculated. P < 0.05 was considered significant.

Data and Resource Availability

No data sets or applicable resources were generated or analyzed during the current study.

Hypothalamic Foxi2 Expression Is Regulated by Nutritional Status

We have previously demonstrated that Foxp1 and Foxq1, two members of the Fox family, regulate hepatic glucose homeostasis (22,23). To further identify potential Fox genes involved in energy metabolism, we next systematically studied the expression of Fox family members in the hypothalamus and other metabolically active tissues under different nutritional statuses by real-time PCR.

Our quantitative PCR and Western blot analyses showed that Foxi2 expression was increased in the hypothalamus of mice with HFD-induced obesity compared with chow-fed control mice (Fig. 1A and C). Furthermore, fasting increased Foxi2 mRNA and protein levels, while refeeding reversed this effect (Fig. 1B and D). Interestingly, the expression patterns of AgRP and NPY in the hypothalamus were similar to that of Foxi2 (Fig. 1E). Of note, Foxi2 was highly expressed in the hypothalamus and rarely expressed in other tissues (Fig. 1F and G). In addition to hypothalamic ARC, Foxi2 was expressed in other brain areas, including the cerebellum and brain stem (Fig. 1H). These data imply that hypothalamic Foxi2 might be involved in systemic energy metabolism.

Figure 1

Hypothalamic Foxi2 expression is regulated by nutritional status. A and C: Quantitative PCR and Western blot analysis of hypothalamic Foxi2 levels of WT mice fed a chow diet or an HFD (n = 6/group). B and D: Quantitative PCR and Western blot analysis of the expression of Foxi2 in the hypothalamus of WT mice under ad libitum (ad-lib)–fed, 36-h–fasted, or 6-h–refed conditions (n = 6/group). E: Quantitative PCR of the expression of AgRP, NPY, POMC, and CART in the hypothalamus of WT mice under ad-lib–fed, 36-h–fasted, or 6-h–refed conditions (n = 6/group). F and G: Male C57BL/6 mice fasted for 6 h were sacrificed for tissue collection. Quantitative PCR and Western blot analysis of Foxi2 expression in the indicated tissues of WT mice (n = 6/group). H: Western blot analysis of Foxi2 protein levels in different brain areas. Data are mean ± SEM. *P < 0.05, **P < 0.01, ***P < 0.001, ****P < 0.0001 by two-tailed Student t test (A) or one-way ANOVA (B and E). PVN, paraventricular nucleus; VMH, ventromedial hypothalamus.

Figure 1

Hypothalamic Foxi2 expression is regulated by nutritional status. A and C: Quantitative PCR and Western blot analysis of hypothalamic Foxi2 levels of WT mice fed a chow diet or an HFD (n = 6/group). B and D: Quantitative PCR and Western blot analysis of the expression of Foxi2 in the hypothalamus of WT mice under ad libitum (ad-lib)–fed, 36-h–fasted, or 6-h–refed conditions (n = 6/group). E: Quantitative PCR of the expression of AgRP, NPY, POMC, and CART in the hypothalamus of WT mice under ad-lib–fed, 36-h–fasted, or 6-h–refed conditions (n = 6/group). F and G: Male C57BL/6 mice fasted for 6 h were sacrificed for tissue collection. Quantitative PCR and Western blot analysis of Foxi2 expression in the indicated tissues of WT mice (n = 6/group). H: Western blot analysis of Foxi2 protein levels in different brain areas. Data are mean ± SEM. *P < 0.05, **P < 0.01, ***P < 0.001, ****P < 0.0001 by two-tailed Student t test (A) or one-way ANOVA (B and E). PVN, paraventricular nucleus; VMH, ventromedial hypothalamus.

Close modal

Foxi2 Activates AgRP Gene Expression by Directly Binding to Its Promoter

Given the similar expression patterns of Foxi2, AgRP, and NPY, we further determined whether Foxi2 is expressed in AgRP neurons. Indeed, our data suggest that Foxi2 is localized in AgRP neurons, as revealed by immunofluorescence (Fig. 2A).

Figure 2

Foxi2 activates AgRP gene expression through directly binding to its promoter. A: Immunofluorescence staining of Foxi2 (green), AgRP (red), and merged (yellow) colocalization in the ARC of WT male mice. Scale bars: 10X, 100 μm; 20X, 50 μm. B: Immunofluorescence staining of Foxi2 (green), AgRP (red), and merged (yellow) colocalization in the ARC of AgRP-IRES-Cre male mice 10 weeks after viral injections. Scale bar: 100 μm. C: Gene expression of Foxi2 in different tissues (ARC, PVN, VMH, cortex, cerebellum, brain stem, liver, WAT, muscle) (n = 6/group). D: Quantitative PCR of hypothalamic AgRP expression in 24-h–fasted AgRP-CreGFP and AgRP-CreFoxi2 mice fed a chow diet (n = 6/group). E: Quantitative PCR of neuropeptide expression in GT1-7 cells treated with Ad-GFP or Ad-Foxi2 for 48 h (n = 3/group). F: Quantitative PCR of mRNA levels of neuropeptides in GT1-7 cells treated with lenti-shGFP or lenti-shFoxi2 for 72 h (n = 3/group). G: Luciferase activity in HEK293 cells transfected with the indicated plasmids. A series of truncated AgRP promoters fused to the luciferase reporter gene were cotransfected into HEK293 cells, together with pcDNA3.1 (control, black bars) or Foxi2 expression plasmids (red bars) (n = 3/group). H: ChIP assay was performed in GT1-7 cells infected with adenovirus expressing Flag-tagged Foxi2 to assess Foxi2 occupancy of the AgRP promoter. The AgRP promoter fragments containing the −345- to −195-bp region could be amplified from the precipitates obtained with an anti-Flag antibody but not with normal mouse IgG (control) (n = 4/group). Data are mean ± SEM. *P < 0.05, **P < 0.01, ***P < 0.001 by two-tailed Student t test (CG) or one-way ANOVA (H). RLA, relative luciferase activity.

Figure 2

Foxi2 activates AgRP gene expression through directly binding to its promoter. A: Immunofluorescence staining of Foxi2 (green), AgRP (red), and merged (yellow) colocalization in the ARC of WT male mice. Scale bars: 10X, 100 μm; 20X, 50 μm. B: Immunofluorescence staining of Foxi2 (green), AgRP (red), and merged (yellow) colocalization in the ARC of AgRP-IRES-Cre male mice 10 weeks after viral injections. Scale bar: 100 μm. C: Gene expression of Foxi2 in different tissues (ARC, PVN, VMH, cortex, cerebellum, brain stem, liver, WAT, muscle) (n = 6/group). D: Quantitative PCR of hypothalamic AgRP expression in 24-h–fasted AgRP-CreGFP and AgRP-CreFoxi2 mice fed a chow diet (n = 6/group). E: Quantitative PCR of neuropeptide expression in GT1-7 cells treated with Ad-GFP or Ad-Foxi2 for 48 h (n = 3/group). F: Quantitative PCR of mRNA levels of neuropeptides in GT1-7 cells treated with lenti-shGFP or lenti-shFoxi2 for 72 h (n = 3/group). G: Luciferase activity in HEK293 cells transfected with the indicated plasmids. A series of truncated AgRP promoters fused to the luciferase reporter gene were cotransfected into HEK293 cells, together with pcDNA3.1 (control, black bars) or Foxi2 expression plasmids (red bars) (n = 3/group). H: ChIP assay was performed in GT1-7 cells infected with adenovirus expressing Flag-tagged Foxi2 to assess Foxi2 occupancy of the AgRP promoter. The AgRP promoter fragments containing the −345- to −195-bp region could be amplified from the precipitates obtained with an anti-Flag antibody but not with normal mouse IgG (control) (n = 4/group). Data are mean ± SEM. *P < 0.05, **P < 0.01, ***P < 0.001 by two-tailed Student t test (CG) or one-way ANOVA (H). RLA, relative luciferase activity.

Close modal

We next determined whether Foxi2 affects the expression of the NPY and AgRP genes. AAV vectors expressing Cre-dependent Foxi2 and GFP as a control (AAV-Foxi2 and AAV-GFP, respectively) were bilaterally injected into the ARC of male AgRP-IRES-Cre mice (27) (Supplementary Fig. 1A). Immunofluorescence staining of Foxi2 and AgRP showed that Foxi2 expression was increased in AgRP-Cre mice and that the majority of Foxi2 was localized in AgRP neurons in the mouse ARC (Fig. 2B). In addition, real-time PCR demonstrated that Foxi2 mRNA levels in the ARC were significantly increased in AAV-Foxi2–infected AgRP-Cre mice (henceforth referred to as AgRP-CreFoxi2 mice) compared with AgRP-CreGFP mice; however, AAV-Foxi2 injection did not alter Foxi2 expression in other brain areas or tissues in the mice (Fig. 2C).

Next, we investigated whether AAV-Foxi2 infection influences the expression of genes encoding AgRP, NPY, POMC, and CART, which are expressed in neurons of the ARC in fasted mice. Our data suggest that AAV-Foxi2 infection increased AgRP gene expression in the ARC of mice but did not significantly alter the expression of other hypothalamic neuropeptides (Fig. 2D). These observations indicate that Foxi2 overexpression specifically increased AgRP expression in vivo.

We also studied the stimulatory effects of Foxi2 on AgRP expression in vitro. We generated a recombinant adenovirus expressing Foxi2 (Ad-Foxi2) and infected hypothalamic GT1-7 cells with this virus. Adenovirus-mediated overexpression of Foxi2 significantly increased the expression of AgRP but did not significantly change the expression of POMC (Fig. 2E). In contrast, Foxi2 knockdown in GT1-7 cells by lenti-shFoxi2 modestly decreased expression of AgRP (Fig. 2F). The above results suggest that Foxi2 is a specific regulator of the neuropeptide AgRP both in vivo and in vitro.

Because Foxi2 belongs to the forkhead family of transcription factors, we wondered whether Foxi2 directly activates AgRP gene transcription. Our luciferase reporter gene assay revealed that overexpression of Foxi2 activated the transcription of pGL3-AgRP promoter regions −1795, −545, and −345 bp, while the stimulatory effects of Foxi2 were abolished when the promoter region was further truncated to −195 bp, suggesting that a Foxi2 binding site was present between −345 and −195 bp (Fig. 2G). Additionally, the occupancy of Foxi2 proteins on the AgRP promoter was also confirmed by ChIP (Fig. 2H). Taken together, these results suggest that Foxi2 is an important regulator of nutrient sensing and can activate AgRP gene expression.

Selective Overexpression of Foxi2 in AgRP Neurons Increases Food Intake and Reduces Energy Expenditure, Subsequently Leading to Obesity

We next explored whether selective Foxi2 overexpression in AgRP neurons of male mice affects energy homeostasis. We observed that the body weight of chow-fed AgRP-CreFoxi2 mice was higher than that of control mice 9 weeks after AAV injection (Fig. 3A and B). MRI examination confirmed that AgRP-CreFoxi2 mice had more fat mass than control mice but no difference in lean mass (Fig. 3C). Furthermore, the individual fat pads of the AgRP-CreFoxi2 mice were markedly larger and weighed more than those of the control mice (Fig. 3D and Supplementary Fig. 1B). Consistently, histological examination (H-E staining) of three fat depots revealed an increase in the size of lipid droplets and adipocytes in the fat pads of AgRP-CreFoxi2 mice (Fig. 3E). However, selective overexpression of Foxi2 did not markedly affect the weight of the kidney, heart, or spleen with a chow diet (data not shown).

Figure 3

Overexpression of Foxi2 in AgRP neurons predisposes mice to obesity, increases food intake, and reduces energy expenditure with a chow diet. A and B: The growth curve starting at 4 weeks after receiving bilateral viral injection into the ARC of 12-week-old male AgRP-IRES-Cre mice and a representative photograph of AgRP-CreGFP and AgRP-CreFoxi2 mice at 23 weeks of age after viral injection (n = 6/group). C: Fat and lean mass (n = 6/group). D: Gross appearance of interscapular BAT and inguinal and epididymal fat pads from the mice in B. E: H-E staining of paraffin-embedded BAT, iWAT, and eWAT sections from AgRP-CreGFP and AgRP-CreFoxi2 mice. Scale bar: 50 μm. F: Daily food intake (n = 6/group). G: Cumulative 24-h food intake in mice fasted overnight (n = 6/group). H and I: Energy expenditure (EE) was measured and analyzed by ANCOVA (n = 6/group). All EE data are expressed as the absolute value (H). All ANCOVA values are expressed as estimated marginal mean ± SEM. The estimated effect size of covariates for ANCOVA is presented as partial η2 (I). J: RQ (VCO2 / VO2) (n = 6/group). K: The rectal temperature of AgRP-CreGFP and AgRP-CreFoxi2 mice subjected to acute cold exposure for 6 h (n = 6/group). L: Quantitative PCR of mRNA levels of genes involved in thermogenesis and fatty acid oxidation in BAT from AgRP-CreGFP and AgRP-CreFoxi2 mice (n = 6/group). M: Western blot analysis of UCP1 protein level in BAT of AgRP-CreGFP and AgRP-CreFoxi2 mice. N and O: Plots for GTTs and ITTs (inset graphs represent the area under the curve [AUC]), respectively, from AgRP-CreGFP and AgRP-CreFoxi2 mice on a chow diet (n = 6/group). P: Quantitative PCR of mRNA levels of UCP1 and Cidea in iWAT of AgRP-CreGFP and AgRP-CreFoxi2 mice subjected to 48-h cold exposure (n = 6/group). Q: Representative images of UCP1 immunohistochemistry of iWAT in AgRP-CreGFP and AgRP-CreFoxi2 mice subjected to cold exposure. Scale bar: 50 μm. R: Western blot analysis of tyrosine hydroxylase (TH) protein level in iWAT from both mice subjected to cold exposure. Data are mean ± SEM. *P < 0.05, **P < 0.01, ***P < 0.001, ****P < 0.0001 by two-tailed Student t test (A, C, FL, N, and O), one-way ANOVA (P), or ANCOVA (I). RT, room temperature.

Figure 3

Overexpression of Foxi2 in AgRP neurons predisposes mice to obesity, increases food intake, and reduces energy expenditure with a chow diet. A and B: The growth curve starting at 4 weeks after receiving bilateral viral injection into the ARC of 12-week-old male AgRP-IRES-Cre mice and a representative photograph of AgRP-CreGFP and AgRP-CreFoxi2 mice at 23 weeks of age after viral injection (n = 6/group). C: Fat and lean mass (n = 6/group). D: Gross appearance of interscapular BAT and inguinal and epididymal fat pads from the mice in B. E: H-E staining of paraffin-embedded BAT, iWAT, and eWAT sections from AgRP-CreGFP and AgRP-CreFoxi2 mice. Scale bar: 50 μm. F: Daily food intake (n = 6/group). G: Cumulative 24-h food intake in mice fasted overnight (n = 6/group). H and I: Energy expenditure (EE) was measured and analyzed by ANCOVA (n = 6/group). All EE data are expressed as the absolute value (H). All ANCOVA values are expressed as estimated marginal mean ± SEM. The estimated effect size of covariates for ANCOVA is presented as partial η2 (I). J: RQ (VCO2 / VO2) (n = 6/group). K: The rectal temperature of AgRP-CreGFP and AgRP-CreFoxi2 mice subjected to acute cold exposure for 6 h (n = 6/group). L: Quantitative PCR of mRNA levels of genes involved in thermogenesis and fatty acid oxidation in BAT from AgRP-CreGFP and AgRP-CreFoxi2 mice (n = 6/group). M: Western blot analysis of UCP1 protein level in BAT of AgRP-CreGFP and AgRP-CreFoxi2 mice. N and O: Plots for GTTs and ITTs (inset graphs represent the area under the curve [AUC]), respectively, from AgRP-CreGFP and AgRP-CreFoxi2 mice on a chow diet (n = 6/group). P: Quantitative PCR of mRNA levels of UCP1 and Cidea in iWAT of AgRP-CreGFP and AgRP-CreFoxi2 mice subjected to 48-h cold exposure (n = 6/group). Q: Representative images of UCP1 immunohistochemistry of iWAT in AgRP-CreGFP and AgRP-CreFoxi2 mice subjected to cold exposure. Scale bar: 50 μm. R: Western blot analysis of tyrosine hydroxylase (TH) protein level in iWAT from both mice subjected to cold exposure. Data are mean ± SEM. *P < 0.05, **P < 0.01, ***P < 0.001, ****P < 0.0001 by two-tailed Student t test (A, C, FL, N, and O), one-way ANOVA (P), or ANCOVA (I). RT, room temperature.

Close modal

We next sought to identify the potential mechanisms of Foxi2 overexpression–induced obesity. We measured food intake and energy expenditure, which can affect energy homeostasis (28). As expected, food intake under an ad libitum feeding condition or after fasting was increased in AgRP-CreFoxi2 mice compared with control mice (Fig. 3F and G), which is consistent with the increased AgRP expression in AgRP-CreFoxi2 mice. In addition, energy consumption and activity levels were monitored by housing mice in metabolic cages. ANCOVA (29) indicated that Foxi2 overexpression significantly influenced energy expenditure during both the day and night cycles (Fig. 3H and I). However, Foxi2 overexpression did not change the utilization of carbohydrates and fat as energy sources, as revealed by RQ data (Fig. 3J). No difference in locomotor activity was observed between AgRP-CreGFP and AgRP-CreFoxi2 mice (Supplementary Fig. 1C). Upon exposure to acute cold, the rectal temperature was lower in AgRP-CreFoxi2 mice (Fig. 3K), further confirming that Foxi2 regulates systemic energy metabolism.

We therefore examined the expression of genes associated with thermogenesis and fatty acid oxidation in BAT harvested from AgRP-CreFoxi2 mice and controls. Consistent with the reduction in metabolic function in AgRP-CreFoxi2 mice, Foxi2 overexpression in AgRP neurons led to a decrease in the expression of thermogenic genes, including UCP1, Dio2, and Cidea. The expression of peroxisome proliferator–activated receptor-α (PPARα), an essential factor involved in fatty acid oxidation, and its target genes, including CPT1a and MCAD, was also decreased in the BAT of AgRP-CreFoxi2 mice (Fig. 3L). Consistently, the protein expression of UCP1 was reduced in AgRP-CreFoxi2 mice (Fig. 3M). We also examined glucose metabolism in these mice. GTTs and ITTs revealed impaired glucose tolerance and insulin resistance in Foxi2-overexpressing mice (Fig. 3N and O). Thus, our results suggest that increased food intake and reduced energy expenditure together contribute to obesity in AgRP-CreFoxi2 mice.

In addition to BAT, beige adipocytes in WAT also contribute to adipose tissue thermogenesis. We first examined the expression of thermogenic genes in the iWAT of AgRP-CreFoxi2 mice housed at room temperature and found that selective Foxi2 overexpression did not influence the expression of thermogenic genes (data not shown). Then, we housed the mice at 4°C for 2 days. Acute cold exposure markedly induced browning of WAT, leading to the appearance of smaller, multilocular beige adipocytes with intense UCP1 immunoreactivity in the iWAT of control mice; however, this effect was markedly attenuated in the iWAT of AgRP-CreFoxi2 mice. Correspondingly, cold exposure induced the expression of genes involved in thermogenesis in the iWAT of AgRP-CreFoxi2 mice, while Foxi2 overexpression in AgRP neurons abolished the stimulatory effects of cold exposure (Fig. 3P and Q). Protein levels of tyrosine hydroxylase, an indicator of sympathetic nerve excitability, were also significantly reduced in the iWAT of AgRP-CreFoxi2 mice following cold exposure (Fig. 3R). These findings suggest that Foxi2 overexpression in AgRP neurons impairs the browning of WAT in response to cold exposure.

Likewise, we obtained similar results in female AgRP-CreFoxi2 and control mice (Supplementary Fig. 2). Thus, Foxi2 regulation of energy metabolism is independent of sex differences.

Furthermore, we studied energy metabolism of these mice under an HFD condition. We found that overexpression of Foxi2 in AgRP neurons aggravates HFD-induced metabolic dysfunction (Supplementary Fig. 3). Compared with control mice, HFD-fed AgRP-CreFoxi2 mice gained more weight because of increased fat mass (Supplementary Fig. 3AF). Likewise, AgRP-CreFoxi2 mice showed an increase in food intake (Supplementary Fig. 3G and H) and decreased energy consumption (Supplementary Fig. 3IL). Molecular mechanism studies indicated that Foxi2 overexpression decreased the expression of genes involved in thermogenesis in BAT and reduced the expression of key components of lipolytic pathways in WAT (Supplementary Fig. 3MO). Furthermore, Foxi2 overexpression aggravated systemic glucose intolerance and impaired insulin sensitivity of metabolic tissues, including liver, WAT, and muscle (Supplementary Fig. 3PT). Finally, compared with controls, AgRP-CreFoxi2 mice showed an aggravated fatty liver phenotype (Supplementary Fig. 3UX).

Foxi2 Overexpression Activates AgRP Neuronal Activity

We next determined whether Foxi2 overexpression influences the activity of AgRP neurons. We measured the activity of AgRP neurons in AgRP-CreGFP and AgRP-CreFoxi2 mice by performing whole-cell patch-clamp recordings on EGFP-labeled AgRP neurons (Fig. 4A).

Figure 4

Increased AgRP neuronal activity upon Foxi2 overexpression. A: A representative image of the AAV-infected AgRP neuron expressing EGFP from AgRP-CreGFP or AgRP-CreFoxi2 mice. The same neuron under bright-field illumination. The merged image of the targeted AgRP neuron is shown in the right panel. B: Resting membrane potential by whole-cell patch-clamp recording in AgRP neurons from fasted AgRP-CreGFP and AgRP-CreFoxi2 mice (n = 6/group). C: Representative traces of APs by patch-clamp recording with injecting current (−10 to 40 pA, step 10 pA) into AgRP neurons from fasted AgRP-CreGFP and AgRP-CreFoxi2 mice. D: The quantification of APs (n = 6/group). E: AP threshold (n = 6/group). F: Representative AP waveforms recorded in AgRP neurons from AgRP-CreGFP and AgRP-CreFoxi2 mice with injection of 20 pA current. G and H: AHP and AP amplitude (n = 6/group). I: Representative traces of sEPSCs in AgRP neurons from fasted AgRP-CreGFP and AgRP-CreFoxi2 mice. J and K: The average frequency and amplitude of sEPSCs recorded in AgRP neurons from both mice (n = 6/group). Data are mean ± SEM. *P < 0.05, **P < 0.01 by two-tailed Student t test (B, D, E, G, H, J, and K).

Figure 4

Increased AgRP neuronal activity upon Foxi2 overexpression. A: A representative image of the AAV-infected AgRP neuron expressing EGFP from AgRP-CreGFP or AgRP-CreFoxi2 mice. The same neuron under bright-field illumination. The merged image of the targeted AgRP neuron is shown in the right panel. B: Resting membrane potential by whole-cell patch-clamp recording in AgRP neurons from fasted AgRP-CreGFP and AgRP-CreFoxi2 mice (n = 6/group). C: Representative traces of APs by patch-clamp recording with injecting current (−10 to 40 pA, step 10 pA) into AgRP neurons from fasted AgRP-CreGFP and AgRP-CreFoxi2 mice. D: The quantification of APs (n = 6/group). E: AP threshold (n = 6/group). F: Representative AP waveforms recorded in AgRP neurons from AgRP-CreGFP and AgRP-CreFoxi2 mice with injection of 20 pA current. G and H: AHP and AP amplitude (n = 6/group). I: Representative traces of sEPSCs in AgRP neurons from fasted AgRP-CreGFP and AgRP-CreFoxi2 mice. J and K: The average frequency and amplitude of sEPSCs recorded in AgRP neurons from both mice (n = 6/group). Data are mean ± SEM. *P < 0.05, **P < 0.01 by two-tailed Student t test (B, D, E, G, H, J, and K).

Close modal

AgRP neurons from 24-h fasted AgRP-CreFoxi2 mice exhibited a depolarized resting membrane potential (Fig. 4B). In addition, the AP firing rate of AgRP neurons evoked by current injection was higher in AgRP-CreFoxi2 mice than in control mice (Fig. 4C and D). Moreover, Foxi2 overexpression in AgRP neurons decreased the AP threshold (Fig. 4E). Furthermore, we analyzed whether selective Foxi2 overexpression affects the shape of APs. After-hyperpolarization (AHP), a component of the AP waveform, is one principal feedback mechanism in controlling the frequency and pattern of neuronal firing. The AHP in AgRP neurons was reduced in AgRP-CreFoxi2 mice, while Foxi2 overexpression in AgRP neurons had no significant effect on the AP amplitude (Fig. 4F–H). Changes in synaptic inputs could affect the spontaneous firing of AgRP neurons. Here, we also examined synaptic transmission at excitatory synapses of AgRP neurons in both groups of mice and found that overexpression of Foxi2 significantly increased the frequency but not the amplitude of sEPSCs in AgRP neurons (Fig. 4I–K). These results suggest that overexpression of Foxi2 increases the excitability of AgRP neurons.

Global Foxi2 Knockout Mice Exhibit Decreased Food Intake and Increased Energy Expenditure

To further examine the physiological function of Foxi2 in energy metabolism, we used global Foxi2 knockout (KO) mice. Western blot analysis failed to detect Foxi2 protein in the hypothalamus of KO mice (Fig. 5A). Compared with WT mice, Foxi2 KO mice showed a reduced mRNA level of AgRP, but the expression levels of NPY, POMC, and CART were not significantly changed (Fig. 5B). Consistently, the protein level of hypothalamic AgRP was reduced in Foxi2 KO mice, as revealed by immunofluorescence staining (Fig. 5C). Indeed, Foxi2 KO mice showed decreased food intake (Fig. 5D and E). The body weight of Foxi2 KO mice was lower than that of control mice starting from 18 weeks of age (Fig. 5F and G). MRI confirmed that Foxi2-deficient mice had less fat mass than control mice, while lean mass remained unchanged (Fig. 5H and I). The individual fat pads of Foxi2-deficient mice, including BAT, iWAT, and eWAT, weighed less than those of control mice (Fig. 5J and Supplementary Fig. 4A). H-E staining revealed that Foxi2 deficiency decreased the size of brown and white adipocytes and lipid droplets in adipocytes (Fig. 5K). Metabolic cage experiments showed that Foxi2 KO mice had increased energy consumption calculated by ANCOVA, whereas there were few differences in RQ or locomotor activity (Fig. 5L–N and Supplementary Fig. 4B). Consistent with the enhanced metabolic function of Foxi2-deficient mice, Foxi2 deficiency led to an increase in the expression of genes involved in thermogenesis and fatty acid oxidation (Fig. 5O and P).

Figure 5

Global Foxi2 KO mice are lean and have decreased food intake and increased energy expenditure with age. A: Western blot analysis of the expression of Foxi2 in the hypothalamus of WT and KO mice. B: Quantitative PCR of hypothalamic neuropeptide expression in 24-h–fasted WT and Foxi2 KO mice fed a chow diet (n = 6/group). C: Immunofluorescence to detect AgRP (red) in WT and KO mice subjected to fasting of 24 h. All nuclei were stained with DAPI (blue). Scale bar: 100 μm. White arrows indicate the third ventricle. D: Daily food intake (n = 6/group). E: Cumulative 24-h food intake in mice fasted overnight (n = 6/group). F and G: The growth curve and a representative photograph of WT and KO mice on a chow diet (n = 6/group). H and I: MRI assay of body composition of WT and KO mice (n = 6/group). J: Gross appearance of interscapular BAT and inguinal and epididymal fat pads from mice in G. K: H-E staining of paraffin-embedded BAT, iWAT, and eWAT sections from WT and KO mice. Scale bar: 50 μm. L and M: Energy expenditure (EE) was measured and analyzed by ANCOVA (n = 6/group). All EE data are expressed as the absolute value (L). All ANCOVA values are expressed as estimated marginal mean ± SEM. Estimated effect size of covariates for ANCOVA is presented as partial η2 (M). N: RQ (VCO2 / VO2) (n = 6/group). O: Quantitative PCR of mRNA levels of genes involved in thermogenesis and fatty acid oxidation in BAT (n = 6/group). P: Western blot analysis of UCP1 protein level in BAT of WT and KO mice. Q and R: Plots for GTTs and ITTs (inset graphs represent the area under the curve [AUC]), respectively, from WT and KO mice fed a chow diet (n = 6/group). Data are mean ± SEM. *P < 0.05, **P < 0.01, ***P < 0.001 by two-tailed Student t test (B, DF, H, I, LO, Q, R) or ANCOVA (M).

Figure 5

Global Foxi2 KO mice are lean and have decreased food intake and increased energy expenditure with age. A: Western blot analysis of the expression of Foxi2 in the hypothalamus of WT and KO mice. B: Quantitative PCR of hypothalamic neuropeptide expression in 24-h–fasted WT and Foxi2 KO mice fed a chow diet (n = 6/group). C: Immunofluorescence to detect AgRP (red) in WT and KO mice subjected to fasting of 24 h. All nuclei were stained with DAPI (blue). Scale bar: 100 μm. White arrows indicate the third ventricle. D: Daily food intake (n = 6/group). E: Cumulative 24-h food intake in mice fasted overnight (n = 6/group). F and G: The growth curve and a representative photograph of WT and KO mice on a chow diet (n = 6/group). H and I: MRI assay of body composition of WT and KO mice (n = 6/group). J: Gross appearance of interscapular BAT and inguinal and epididymal fat pads from mice in G. K: H-E staining of paraffin-embedded BAT, iWAT, and eWAT sections from WT and KO mice. Scale bar: 50 μm. L and M: Energy expenditure (EE) was measured and analyzed by ANCOVA (n = 6/group). All EE data are expressed as the absolute value (L). All ANCOVA values are expressed as estimated marginal mean ± SEM. Estimated effect size of covariates for ANCOVA is presented as partial η2 (M). N: RQ (VCO2 / VO2) (n = 6/group). O: Quantitative PCR of mRNA levels of genes involved in thermogenesis and fatty acid oxidation in BAT (n = 6/group). P: Western blot analysis of UCP1 protein level in BAT of WT and KO mice. Q and R: Plots for GTTs and ITTs (inset graphs represent the area under the curve [AUC]), respectively, from WT and KO mice fed a chow diet (n = 6/group). Data are mean ± SEM. *P < 0.05, **P < 0.01, ***P < 0.001 by two-tailed Student t test (B, DF, H, I, LO, Q, R) or ANCOVA (M).

Close modal

We also examined glucose metabolism in these mice. GTTs and ITTs revealed that Foxi2 KO mice had enhanced glucose tolerance and improved insulin sensitivity (Fig. 5Q and R). These results suggest that depletion of Foxi2 decreases food intake and increases energy expenditure, subsequently improving metabolic phenotypes.

Deficiency of Foxi2 Alleviates HFD-Induced Metabolic Dysfunction

We also challenged Foxi2-deficient mice with an HFD and studied the metabolic phenotype of these mice. We fed 6-week-old male WT and Foxi2 KO mice an HFD for 16 weeks and obtained similar results as those reported above. Foxi2 KO mice showed a decrease in food intake (Fig. 6A and B). After 16 weeks of HFD feeding, Foxi2 KO mice displayed reduced body weight, fat mass, and BAT, iWAT, and eWAT weight (Fig. 6C–F and Supplementary Fig. 5A). Likewise, H-E staining suggested that the size of adipocytes in three different fat depots in Foxi2 KO mice was also reduced (Fig. 6G). HFD-fed Foxi2 KO mice exhibited increased energy consumption and decreased RQ but no change in locomotor activity (Fig. 6H–J and Supplementary Fig. 5B). Furthermore, Foxi2 KO mice had an increased rectal temperature, confirming the enhancement of energy expenditure (Fig. 6K). Likewise, the expression levels of thermogenic genes in BAT were increased in Foxi2 KO mice (Fig. 6L). Additionally, Foxi2 deficiency increased the protein levels of p-HSL and ATGL, the rate-limiting enzymes involved in lipolysis, in iWAT (Fig. 6M). These results suggest that Foxi2 deficiency can protect against HFD-induced obesity.

Figure 6

Foxi2 KO mice are resistant to HFD-induced obesity. A: Daily food intake (n = 6/group). B: Cumulative food intake of 24 h in WT and KO mice fasted overnight (n = 6/group). C and D: The growth curve and a representative photograph of WT and Foxi2 KO mice at 22 weeks of age fed an HFD (n = 6/group). E: MRI assay of body composition of WT and KO mice (n = 6/group). F: Gross appearance of interscapular BAT and inguinal and epididymal fat pads from mice in D. G: H-E staining of paraffin-embedded BAT, iWAT, and eWAT sections from WT and KO mice. Scale bar: 50 μm. H and I: Energy expenditure (EE) was measured and analyzed by ANCOVA (n = 6/group). All EE data are expressed as the absolute value (H). All ANCOVA values are expressed as estimated marginal mean ± SEM. Estimated effect size of covariates for ANCOVA analysis is presented as partial η2 (I). J: RQ (VCO2 / VO2) (n = 6/group). K: The rectal temperature of WT and KO mice subjected to acute cold exposure for 6 h (n = 6/group). L: Quantitative PCR of mRNA levels of UCP1, PGC-1α, Cidea, and Nrf1 in BAT from WT and KO mice (n = 6/group). M: Western blot analysis of lipolytic enzymes in eWAT from WT and KO mice. N and O: Fed blood glucose and insulin levels in WT and KO mice (n = 6/group). P and Q: Plots for GTTs and ITTs (inset graphs represent the area under the curve [AUC]), respectively, from WT and KO mice on an HFD (n = 6/group). R: Western blot analysis of AKT and p-AKT (Ser473) in the liver, adipose tissue, and muscle 15 min after administration of insulin (1 unit/kg). S: Representative photograph of liver in WT and KO mice. T: Representative images of H-E– and oil red O–stained hepatic sections of WT and KO mice. Scale bar: 100 μm. U: Hepatic TG and TC content and serum TG and TC content of WT and KO mice (n = 6/group). V: Quantitative PCR of Srebp-1c, fatty acid synthase (Fasn), acetyl CoA carboxylase (Acc), stearoyl-CoA desaturase 1 (Scd1), CPT1a, PPARα, and PGC-1α in the livers of WT and KO mice (n = 6/group). Data are mean ± SEM. *P < 0.05, **P < 0.01, ***P < 0.001 by two-tailed Student t test (AC, E, HL, NQ, U, V) or ANCOVA (I).

Figure 6

Foxi2 KO mice are resistant to HFD-induced obesity. A: Daily food intake (n = 6/group). B: Cumulative food intake of 24 h in WT and KO mice fasted overnight (n = 6/group). C and D: The growth curve and a representative photograph of WT and Foxi2 KO mice at 22 weeks of age fed an HFD (n = 6/group). E: MRI assay of body composition of WT and KO mice (n = 6/group). F: Gross appearance of interscapular BAT and inguinal and epididymal fat pads from mice in D. G: H-E staining of paraffin-embedded BAT, iWAT, and eWAT sections from WT and KO mice. Scale bar: 50 μm. H and I: Energy expenditure (EE) was measured and analyzed by ANCOVA (n = 6/group). All EE data are expressed as the absolute value (H). All ANCOVA values are expressed as estimated marginal mean ± SEM. Estimated effect size of covariates for ANCOVA analysis is presented as partial η2 (I). J: RQ (VCO2 / VO2) (n = 6/group). K: The rectal temperature of WT and KO mice subjected to acute cold exposure for 6 h (n = 6/group). L: Quantitative PCR of mRNA levels of UCP1, PGC-1α, Cidea, and Nrf1 in BAT from WT and KO mice (n = 6/group). M: Western blot analysis of lipolytic enzymes in eWAT from WT and KO mice. N and O: Fed blood glucose and insulin levels in WT and KO mice (n = 6/group). P and Q: Plots for GTTs and ITTs (inset graphs represent the area under the curve [AUC]), respectively, from WT and KO mice on an HFD (n = 6/group). R: Western blot analysis of AKT and p-AKT (Ser473) in the liver, adipose tissue, and muscle 15 min after administration of insulin (1 unit/kg). S: Representative photograph of liver in WT and KO mice. T: Representative images of H-E– and oil red O–stained hepatic sections of WT and KO mice. Scale bar: 100 μm. U: Hepatic TG and TC content and serum TG and TC content of WT and KO mice (n = 6/group). V: Quantitative PCR of Srebp-1c, fatty acid synthase (Fasn), acetyl CoA carboxylase (Acc), stearoyl-CoA desaturase 1 (Scd1), CPT1a, PPARα, and PGC-1α in the livers of WT and KO mice (n = 6/group). Data are mean ± SEM. *P < 0.05, **P < 0.01, ***P < 0.001 by two-tailed Student t test (AC, E, HL, NQ, U, V) or ANCOVA (I).

Close modal

Biochemical analysis revealed that serum glucose and insulin levels were significantly reduced in Foxi2 KO mice fed an HFD (Fig. 6N and O). GTTs and ITTs confirmed that Foxi2 deficiency improved glucose intolerance and insulin resistance induced by HFD feeding (Fig. 6P and Q). Consistent with the ITT data, insulin-stimulated AKT activation was enhanced in Foxi2 KO mice (Fig. 6R). Moreover, Foxi2 KO mice showed reduced liver weight and hepatic lipid accumulation, as revealed by H-E and oil red O staining (Fig. 6S and T and Supplementary Fig. 5C). Hepatic and serum TG levels, as well as serum cholesterol levels, were reduced in HFD-fed Foxi2 KO mice (Fig. 6U). Likewise, Foxi2 deficiency decreased the mRNA levels of lipogenic genes in the livers of mice, while the expression of CPT1a, PPARα, and peroxisome proliferator–activated receptor γ coactivator 1α (PGC-1α) was not markedly changed (Fig. 6V). Collectively, these data indicate that Foxi2 deficiency alleviates hepatic steatosis and improves insulin resistance in HFD-fed mice.

AgRP neurons are activated by hormonal and neural signals of heightened energy demand and stimulate feeding, at least in part, by antagonizing postsynaptic MC3R/MC4Rs through the release of AgRP (30). In the current study, we show that the transcription factor Foxi2, which was highly expressed in the hypothalamus, is a critical regulator of AgRP expression in response to nutritional status. We found that nutritional status, including fasting and HFD feeding, regulates Foxi2 expression. Fasting is well known to induce AgRP gene expression and enhance neuronal activity (31). How fasting induces Foxi2 expression and whether Foxi2 mediates the effects of fasting on AgRP neuron activity deserve further study.

In addition to Foxi2, other factors, including ATF4 and Foxo1, have been shown to regulate AgRP gene expression and neuronal activity (32,33). Of note, ATF4 has been demonstrated to regulate AgRP expression through Foxo1. Additionally, a consensus site for Stat3 exists in the promoter region of the AgRP gene, and overexpression of Stat3 causes inhibition of AgRP transcription (33,34). Thus, multiple factors form a complicated transcriptional network regulating AgRP expression in response to hormonal and nutritional cues. It is logical to speculate that Foxi2, as a transcription factor, has multiple downstream target genes and other physiological roles in the hypothalamus. ChIP sequencing in combination with RNA sequencing is required to identify the global target genes of Foxi2 in the hypothalamus. In addition to the ARC, our data suggest that Foxi2 is also expressed in the paraventricular nucleus, ventromedial hypothalamus, cerebellum, and brain stem. Further studies are required to determine the physiological functions of Foxi2 in these brain areas.

FOX proteins constitute an evolutionarily conserved family of transcription factors with an important role in development and metabolism. Based on sequence conservation, FOX genes have been divided into 19 subfamilies from Foxa to Foxs (35). Among the best studied of these factors are Foxo1 and Foxa2, both of which have been shown to regulate systemic glucose and lipid metabolism (36,37). The activity of Foxo1 is regulated by posttranslational modification, including phosphorylation and acetylation. These modifications affect Foxo1 location from the nucleus to the cytosol (36). Similar results were also observed for Foxa2, which is phosphorylated by insulin/phosphatidylinositol 3-kinase/AKT signaling and acetylated by p300 (38,39). Whether and how Foxi2 proteins are regulated by these posttranslational modifications in response to hormonal and nutritional cues remains unclear. Further studies are required to clarify these questions.

Foxi2 overexpression in AgRP neurons induced AgRP gene expression, thereby increasing food intake and decreasing energy metabolism, and subsequently caused glucose intolerance and impaired insulin sensitivity and obesity. Of note, to explore the effects of Foxi2 overexpression in AgRP neurons, we performed stereotactic injection of AAV-GFP (as a control) and AAV-Foxi2 into the ARC of AgRP-IRES-Cre mice. Each group consisted of seven mice. To verify the anatomical accuracy and effectiveness of viral injection, we examined the Foxi2 expression levels in ARC and other brain areas of all these sacrificed mice by quantitative PCR and immunofluorescence assay. We found that Foxi2 was specifically induced in the ARC of six AAV-Foxi2–injected AgRP-IRES-Cre mice, while its expression was not altered in the ARC of one mouse. Thus, this mouse was excluded from further study.

In contrast, global Foxi2-deficient mice gained less weight upon aging and were resistant to HFD-induced obesity. In the current study, we used male mice in most cases, except we used chow-fed female mice with selective overexpression of Foxi2 in AgRP neurons. This is a weakness of the study. In the future, female Foxi2 KO mice are required to further confirm our results. Additionally, we performed GTTs and ITTs using the traditional methodology, which has several weaknesses. First, in a typical GTT experiment, mice are fasted overnight. However, a shorter fast (5–6 h) is more physiological for mice. Second, the traditional approach for GTTs or ITTs performed in mice is to base the dose of glucose or insulin on the weight of the mouse. Lean mass (muscle, brain, and liver) is the principal site of glucose disposal. However, modulating Foxi2 expression in mice changed the fat mass but not the lean mass. Thus, it is more appropriate to base the dose of glucose and insulin on lean mass rather than on body weight (40). Although ANCOVA showed that modulating Foxi2 expression in mice similarly influences energy consumption and body weight as age, we point out that these mice were maintained at room temperature (∼23°C) rather than at thermoneutrality (30°C). The 23°C house condition is a chronic thermal stress to mice, which may have influenced the outcome of the metabolic studies (41,42). It is more appropriate to study the effects of Foxi2 on energy metabolism of mice at thermoneutrality to avoid the thermal stress.

We show that global Foxi2 KO mice have decreased food intake and increased energy metabolism. Notably, Foxi2 is also expressed in other brain areas and BAT. Thus, Foxi2 deficiency in these tissues may also contribute to these phenotypes of Foxi2 KO mice. Moreover, Foxi2 may be involved in brain development (15,43), which might indirectly affect systemic energy metabolism. Thus, AgRP-specific Foxi2 KO mice were required to confirm its function in AgRP neurons in the future.

Previous studies have indicated that AgRP expression levels are positively related to AgRP neuron activity and food intake. For example, fasting induces AgRP mRNA levels and activates AgRP neuron activity (31). Leptin and insulin act in the ARC of the hypothalamus to reduce expression of the AgRP gene and inhibit AgRP neuron activity (30,33,44). Furthermore, Foxo1 activates AgRP gene transcription and influences food intake of mice in response to leptin treatment (33). Finally, PGC-1α regulates AgRP expression in response to fasting, and PGC-1α deficiency in AgRP neurons led to a reduction of food intake (45). Consistent with these data, our data indicate that Foxi2 overexpression stimulates AgRP expression and enhances AgRP neuronal activity (whole-cell patch-clamp experiments). Thus, it appears that an increase in AgRP expression levels drives the activity of AgRP neurons on the basis of these data. On the other hand, insulin and leptin have been shown to inhibit AgRP neuronal activity through the phosphatidylinositol 3-kinase/phosphatidylinositol 4-phosphate 3 axis. Phosphatidylinositol 4-phosphate 3 opens KATP channels, thereby producing an outward flow of K+ ions and neuron hyperpolarization (46). Whether and how Foxi2 influences K+ ions and AgRP neuron hyperpolarization remain unknown. The exact mechanism of Foxi2 activation of AgRP neuron activity deserves further studies.

In summary, our work identifies Foxi2 as a critical regulator of AgRP neuron activity. Targeting Foxi2 in the hypothalamus might be a therapeutic strategy for treating obesity and metabolic diseases.

This article contains supplementary material online at https://doi.org/10.2337/figshare.20332899.

Y.F., S.S., and C.G. contributed equally.

Funding. This work was supported by the National Natural Science Foundation of China (grants 81825004 and 81730024), the National Key Research and Development Program of China (grant 2018YFA0800601), and the Scientific and Technological Research Project of Xinjiang Production and Construction Corps (grants 2018AB018 and 2021AB028).

Duality of Interest. No potential conflicts of interest relevant to this article were reported.

Author Contributions. Y.F., S.S., and C.G. conducted the experiments. Y.F., S.S., C.G., W.Q., Y.J., L.T., L.Z., and X.D. acquired data. Y.F., S.S., C.G., W.Q., Y.J., Y.G., L.Z., J.Z., and X.D. analyzed data. Y.F., Y.L., H.S., and Y.C. designed the research studies. Y.F. and Y.C. wrote the manuscript. L.T., Y.G., J.Z., X.L., H.S., Y.L, and Y.C. provided reagents. Y.C. is the guarantor of this work and, as such, had full access to all the data in the study and takes responsibility for the integrity of the data and the accuracy of the data analysis.

1.
Haslam
DW
,
James
WP
.
Obesity
.
Lancet
2005
;
366
:
1197
1209
2.
Kolotkin
RL
,
Meter
K
,
Williams
GR
.
Quality of life and obesity
.
Obes Rev
2001
;
2
:
219
229
3.
Spiegelman
BM
,
Flier
JS
.
Obesity and the regulation of energy balance
.
Cell
2001
;
104
:
531
543
4.
Berthoud
HR
,
Münzberg
H
,
Morrison
CD
.
Blaming the brain for obesity: integration of hedonic and homeostatic mechanisms
.
Gastroenterology
2017
;
152
:
1728
1738
5.
Thomas
MA
,
Xue
B
.
Mechanisms for AgRP neuron-mediated regulation of appetitive behaviors in rodents
.
Physiol Behav
2018
;
190
:
34
42
6.
Tong
Q
,
Ye
CP
,
Jones
JE
,
Elmquist
JK
,
Lowell
BB
.
Synaptic release of GABA by AgRP neurons is required for normal regulation of energy balance
.
Nat Neurosci
2008
;
11
:
998
1000
7.
Flier
JS
.
AgRP in energy balance: will the real AgRP please stand up?
Cell Metab
2006
;
3
:
83
85
8.
Williams
KW
,
Elmquist
JK
.
Lighting up the hypothalamus: coordinated control of feeding behavior
.
Nat Neurosci
2011
;
14
:
277
278
9.
Krashes
MJ
,
Koda
S
,
Ye
C
, et al
.
Rapid, reversible activation of AgRP neurons drives feeding behavior in mice
.
J Clin Invest
2011
;
121
:
1424
1428
10.
Joly-Amado
A
,
Denis
RG
,
Castel
J
, et al
.
Hypothalamic AgRP-neurons control peripheral substrate utilization and nutrient partitioning
.
EMBO J
2012
;
31
:
4276
4288
11.
Kaestner
KH
,
Knochel
W
,
Martinez
DE
.
Unified nomenclature for the winged helix/forkhead transcription factors
.
Genes Dev
2000
;
14
:
142
146
12.
Solomon
KS
,
Logsdon
JM
Jr
,
Fritz
A
.
Expression and phylogenetic analyses of three zebrafish FoxI class genes
.
Dev Dyn
2003
;
228
:
301
307
13.
Hulander
M
,
Wurst
W
,
Carlsson
P
,
Enerbäck
S
.
The winged helix transcription factor Fkh10 is required for normal development of the inner ear
.
Nat Genet
1998
;
20
:
374
376
14.
Edlund
RK
,
Ohyama
T
,
Kantarci
H
,
Riley
BB
,
Groves
AK
.
Foxi transcription factors promote pharyngeal arch development by regulating formation of FGF signaling centers
.
Dev Biol
2014
;
390
:
1
13
15.
Wijchers
PJ
,
Hoekman
MF
,
Burbach
JP
,
Smidt
MP
.
Cloning and analysis of the murine Foxi2 transcription factor
.
Biochim Biophys Acta
2005
;
1731
:
133
138
16.
Seale
P
,
Conroe
HM
,
Estall
J
, et al
.
Prdm16 determines the thermogenic program of subcutaneous white adipose tissue in mice
.
J Clin Invest
2011
;
121
:
96
105
17.
Leone
TC
,
Lehman
JJ
,
Finck
BN
, et al
.
PGC-1alpha deficiency causes multi-system energy metabolic derangements: muscle dysfunction, abnormal weight control and hepatic steatosis
.
PLoS Biol
2005
;
3
:
e101
18.
Ortega-Molina
A
,
Efeyan
A
,
Lopez-Guadamillas
E
, et al
.
Pten positively regulates brown adipose function, energy expenditure, and longevity
.
Cell Metab
2012
;
15
:
382
394
19.
Tang
Q
,
Gao
Y
,
Liu
Q
, et al
.
Sirt6 in pro-opiomelanocortin neurons controls energy metabolism by modulating leptin signaling
.
Mol Metab
2020
;
37
:
100994
20.
Patwari
P
,
Emilsson
V
,
Schadt
EE
, et al
.
The arrestin domain-containing 3 protein regulates body mass and energy expenditure
.
Cell Metab
2011
;
14
:
671
683
21.
Yang
L
,
Zhou
ZG
,
Sun
XF
.
Specific knockdown of PPARδ gene in colon cancer cells by lentivirus-mediated RNA interfering
.
Methods Mol Biol
2013
;
952
:
67
86
22.
Zou
Y
,
Gong
N
,
Cui
Y
, et al
.
Forkhead box P1 (FOXP1) transcription factor regulates hepatic glucose homeostasis
.
J Biol Chem
2015
;
290
:
30607
30615
23.
Cui
Y
,
Qiao
A
,
Jiao
T
, et al
.
The hepatic FOXQ1 transcription factor regulates glucose metabolism in mice
.
Diabetologia
2016
;
59
:
2229
2239
24.
Yao
L
,
Cui
X
,
Chen
Q
, et al
.
Cold-inducible SIRT6 regulates thermogenesis of brown and beige fat
.
Cell Rep
2017
;
20
:
641
654
25.
Cui
A
,
Fan
H
,
Zhang
Y
, et al
.
Dexamethasone-induced Krüppel-like factor 9 expression promotes hepatic gluconeogenesis and hyperglycemia
.
J Clin Invest
2019
;
129
:
2266
2278
26.
Fang
X
,
Jiang
S
,
Wang
J
, et al
.
Chronic unpredictable stress induces depression-related behaviors by suppressing AgRP neuron activity
.
Mol Psychiatry
2021
;
26
:
2299
2315
27.
Mandelblat-Cerf
Y
,
Ramesh
RN
,
Burgess
CR
, et al
.
Arcuate hypothalamic AgRP and putative POMC neurons show opposite changes in spiking across multiple timescales
.
eLife
2015
;
4
:
e07122
28.
Morton
GJ
,
Meek
TH
,
Schwartz
MW
.
Neurobiology of food intake in health and disease
.
Nat Rev Neurosci
2014
;
15
:
367
378
29.
Morris
EM
,
Noland
RD
,
Allen
JA
, et al
.
Difference in housing temperature-induced energy expenditure elicits sex-specific diet-induced metabolic adaptations in mice
.
Obesity (Silver Spring)
2020
;
28
:
1922
1931
30.
Deem
JD
,
Faber
CL
,
Morton
GJ
.
AgRP neurons: regulators of feeding, energy expenditure, and behavior
.
FEBS J
2022
;
289
:
2362
2381
31.
Liu
T
,
Kong
D
,
Shah
BP
, et al
.
Fasting activation of AgRP neurons requires NMDA receptors and involves spinogenesis and increased excitatory tone
.
Neuron
2012
;
73
:
511
522
32.
Deng
J
,
Yuan
F
,
Guo
Y
, et al
.
Deletion of ATF4 in AgRP neurons promotes fat loss mainly via increasing energy expenditure
.
Diabetes
2017
;
66
:
640
650
33.
Kitamura
T
,
Feng
Y
,
Kitamura
YI
, et al
.
Forkhead protein FoxO1 mediates Agrp-dependent effects of leptin on food intake
.
Nat Med
2006
;
12
:
534
540
34.
Brown
AM
,
Mayfield
DK
,
Volaufova
J
,
Argyropoulos
G
.
The gene structure and minimal promoter of the human agouti related protein
.
Gene
2001
;
277
:
231
238
35.
Benayoun
BA
,
Caburet
S
,
Veitia
RA
.
Forkhead transcription factors: key players in health and disease
.
Trends Genet
2011
;
27
:
224
232
36.
van der Horst
A
,
Burgering
BM
.
Stressing the role of FoxO proteins in lifespan and disease
.
Nat Rev Mol Cell Biol
2007
;
8
:
440
450
37.
Friedman
JR
,
Kaestner
KH
.
The Foxa family of transcription factors in development and metabolism
.
Cell Mol Life Sci
2006
;
63
:
2317
2328
38.
Wolfrum
C
,
Asilmaz
E
,
Luca
E
,
Friedman
JM
,
Stoffel
M
.
Foxa2 regulates lipid metabolism and ketogenesis in the liver during fasting and in diabetes
.
Nature
2004
;
432
:
1027
1032
39.
von Meyenn
F
,
Porstmann
T
,
Gasser
E
, et al
.
Glucagon-induced acetylation of Foxa2 regulates hepatic lipid metabolism
.
Cell Metab
2013
;
17
:
436
447
40.
Ayala
JE
,
Samuel
VT
,
Morton
GJ
, et al.;
NIH Mouse Metabolic Phenotyping Center Consortium
.
Standard operating procedures for describing and performing metabolic tests of glucose homeostasis in mice
.
Dis Model Mech
2010
;
3
:
525
534
41.
Feldmann
HM
,
Golozoubova
V
,
Cannon
B
,
Nedergaard
J
.
UCP1 ablation induces obesity and abolishes diet-induced thermogenesis in mice exempt from thermal stress by living at thermoneutrality
.
Cell Metab
2009
;
9
:
203
209
42.
Golozoubova
V
,
Gullberg
H
,
Matthias
A
,
Cannon
B
,
Vennström
B
,
Nedergaard
J
.
Depressed thermogenesis but competent brown adipose tissue recruitment in mice devoid of all hormone-binding thyroid hormone receptors
.
Mol Endocrinol
2004
;
18
:
384
401
43.
Ohyama
T
,
Groves
AK
.
Expression of mouse Foxi class genes in early craniofacial development
.
Dev Dyn
2004
;
231
:
640
646
44.
Mayer
CM
,
Belsham
DD
.
Insulin directly regulates NPY and AgRP gene expression via the MAPK MEK/ERK signal transduction pathway in mHypoE-46 hypothalamic neurons
.
Mol Cell Endocrinol
2009
;
307
:
99
108
45.
Gill
JF
,
Delezie
J
,
Santos
G
,
Handschin
C
.
PGC-1α expression in murine AgRP neurons regulates food intake and energy balance
.
Mol Metab
2016
;
5
:
580
588
46.
Könner
AC
,
Klöckener
T
,
Brüning
JC
.
Control of energy homeostasis by insulin and leptin: targeting the arcuate nucleus and beyond
.
Physiol Behav
2009
;
97
:
632
638
Readers may use this article as long as the work is properly cited, the use is educational and not for profit, and the work is not altered. More information is available at https://www.diabetesjournals.org/journals/pages/license.