Skip to main content
eNeuro logoLink to eNeuro
. 2017 Oct 11;4(5):ENEURO.0186-17.2017. doi: 10.1523/ENEURO.0186-17.2017

Luciferase shRNA Presents off-Target Effects on Voltage-Gated Ion Channels in Mouse Hippocampal Pyramidal Neurons

Yuto Hasegawa 1,*, Wenjie Mao 1,*, Sucharita Saha 1, Georgia Gunner 1, Jenya Kolpakova 1, Gilles E Martin 1, Kensuke Futai 1,
PMCID: PMC5635487  PMID: 29034317

Abstract

RNA interference (RNAi) is a straightforward approach to study gene function from the in vitro cellular level to in vivo animal behavior. Although RNAi-mediated gene knockdown has become essentially routine in neuroscience over the past ten years, off-target effects of short hairpin RNAs (shRNAs) should be considered as the proper choice of control shRNA is critical in order to perform meaningful experiments. Luciferase shRNA (shLuc), targeting firefly luciferase, and scrambled shRNAs (shScrs) have been widely used as controls for vertebrate cell research. However, thorough validation of control shRNAs has not been made to date. Here, we performed thorough physiological and morphological studies against control shRNAs in mouse hippocampal CA1 pyramidal neurons. As expected, all control shRNAs exhibited normal basal synaptic transmission and dendritic morphology. However, to our surprise, shLuc exerted severe off-target effects on voltage-gated ion channel function, while the shScr had no detectable changes. These results indicate that thorough validation of shRNA is imperative and, in the absence of such validation, that shScr is the best available negative control for gene knockdown studies.

Keywords: ion channels, luciferase, off-target effects, RNA interference, short hairpin

Significance Statement

RNA interference (RNAi), a process through which small RNAs induce sequence-specific post-transcriptional gene silencing, is widely recognized as one of the most ideal tools not only for functional genomics but also for therapeutic applications. Ensuring the specificity of small interfering RNAs (siRNAs) for specific messenger RNAs is critical, as off-target effects of siRNA can compromise the interpretation of data. Since the existence of off-target effects has been suggested in the past, it is critical to unequivocally establish that siRNAs do not present unintended effects on the biophysical properties of neurons. Here, we found that luciferase short hairpin RNA (shRNA), but not scrambled shRNA (shScr), exhibited off-target effects on voltage-gated ion channels, indicating that careful evaluation is required for studies using luciferase shRNA (shLuc) and siRNA in general.

Introduction

Since the finding of RNA interference (RNAi; Fire et al., 1998), RNAi-mediated gene knockdown has become one of the most straightforward and high throughput approaches to address gene functions from the cellular level to the animal level. Although this approach is technically straightforward, off-target effects have been repeatedly reported and represent a major concern when using RNAi gene knockdown technology. Three off-target mechanisms have been described: (1) microRNA-like regulation through sequence complementary to the small interfering RNA (siRNA) seed region (Jackson et al., 2003; Birmingham et al., 2006; Jackson et al., 2006a,b); (2) Toll-like receptor-mediated immune stimulation (Sledz et al., 2003; Kariko et al., 2004); (3) oversaturation of endogenous RNAi machinery by siRNA and short hairpin RNA (shRNA) transfection (Grimm et al., 2006; Khan et al., 2009). Therefore, it is important to perform multiple layers of confirmatory experiments to validate the effect(s) of RNAi, such as verifying phenotypes by independent multiple siRNAs, rescuing phenotypes by exogenous transgene expression and including nonspecific RNAi control(s) (Cullen, 2006).

shRNAs, the most widely used double-stranded RNAs, are processed to siRNAs by Dicer and silence target genes along the RISC-mediated RNAi pathway. shRNA directed against firefly luciferase (shLuc) has been widely used as a control in mammalian cells. To date, >70 publications used shLuc as a control. However, control shRNAs, including shLuc, have not been fully validated. Alvarez et al. (2006) reported the off-target effects of shLuc in hippocampal CA1 pyramidal neurons. Their findings indicate that transfection of shLuc caused dysregulation of spine density and dendritic complexity, and concomitant reduction of excitatory and inhibitory synaptic transmission. Despite these striking results, shLuc continues to be used as a control (Hoogenraad et al., 2010; Wakita et al., 2011; Chen et al., 2014a,b), indicating the need for a thorough validation of control shRNAs in cellular function.

Here, we have performed a detailed characterization of two control shRNAs on neuronal function and morphology, and demonstrate that shLuc has considerable off-target effects on voltage-gated ion channels without exhibiting any synaptic or morphologic defects. In contrast, nonsilencing scrambled shRNA (shScr) exhibited no abnormal neuronal functions and morphology. This study highlights the importance of thoroughly validating shRNAs and proposes shScr as a negative control appropriate for gene knockdown studies in mammalian cells.

Materials and Methods

Animals

All animal protocols were approved by the Institutional Animal Care and Use Committee of the University of Massachusetts Medical School. Male and female C57BL6 mice were used.

DNA constructs

The human H1 promoter-based pSuper Luciferase-RNAi construct has been previously described (Zhang and Macara, 2006) and targets the sequence 5´-CGTACGCGGAATACTTCGA-3´. pGIPZ-shScr (Dharmacon, #RHS4346) targets the sequence 5´-ATCTCGCTTGGGCGAGAGTAAG-3´. Dharmacon nontargeting scrambled sequence was designed using a proprietary algorithm to ensure the sequence will not target any annotated gene in human, mouse, or rat. EGFP (Clontech) gene was sub-cloned to pCAG vector.

Organotypic slice culture preparation and biolistic gene transfection

Mice hippocampal organotypic slice cultures were prepared from postnatal day 5–7 C57BL6 mice (both genders; Stoppini et al., 1991; Futai et al., 2007; Futai et al., 2013). Briefly, hippocampal slices (350-μm thickness) were prepared using a tissue chopper (Ted Pella, INC), and slices were cultured in a CO2 incubator at 35°C. Neurons were transfected at days in vitro 4–6 using a biolistic gene gun (Bio-Rad; Lo et al., 1994) with 1.6-μm gold particles (10 mg per ∼50 bullets) coated with cDNAs: shRNA vector, pCAG and pCAG-EGFP (45:45:10 μg), and were assayed 5 d after transfection, unless otherwise noted.

Electrophysiology

The recording chamber was filled with extracellular solution containing 119 mM NaCl, 2.5 mM KCl, 4 mM CaCl2, 4 mM MgCl2, 26 mM NaHCO3, 1 mM NaH2PO4, and 11 mM glucose, gassed with 5% CO2/95% O2, pH 7.4. For whole-cell recordings, thick-walled borosilicate glass pipettes (Warner Instruments) were pulled to a resistance of 2–4 MΩ. Na+ currents were measured in the presence of K+ and calcium (Ca2+) channel blockers, tetraethylammonium (TEA; 30 mM, Sigma), 4-amynopyridine (4-AP; 0.5 mM, Sigma), and CdCl2 (100 mM) in extracellular solution. To measure K+ currents, tetrodotoxin (TTX; 1 μM) and CdCl2 were applied to extracellular solution to block Na+ and Ca2+ channels.

Current-clamp recordings were performed with glass electrodes filled with internal solution containing the following: 115 mM potassium methanesulfonate, 20 mM KCl, 10 mM HEPES, 2.5 mM MgCl2, 4 mM adenosine triphosphate disodium salt, 0.4 mM guanosine triphosphate trisodium salt, 10 mM sodium phosphocreatine, and 0.6 mM EGTA, pH 7.25, with KOH. For voltage-clamp recordings, the potassium was replaced by cesium.

All experiments and the analysis of data were performed in a blind manner. Recordings were performed using a MultiClamp 700B amplifier and Digidata 1440, and data were acquired and analyzed using Clampex 10.3 and Clampfit 10.3 (Molecular Devices).

Imaging

The organotypic slices were fixed in 4% paraformaldehyde and 4% sucrose in PBS overnight. The slices were then cryoprotected in 30% sucrose in 0.1 M phosphate buffer (pH 7.4) for 2 h at room temperature followed by rapid freeze and thaw treatments. The slices were then stained with antibodies [γ-H2AX (Millipore, #05-636), GFP (MBL, #598), and secondary Alexa Fluor 647/488 dye-conjugated anti-mouse (Jackson ImmunoResearch) antibodies] in GDB buffer (0.1% gelatin, 0.5% TX-100, 450 mM NaCl, and 32% 0.1 M phosphate buffer, pH 7.4; McAllister, 2000). Two-photon and confocal microscopy were used for spine and γ-H2AX imaging, respectively. Two-photon images were captured using a LUMPlanN 60× (1.0 NA) objective. All secondary dendrites for each neuron were subjected to spine dimension and density analysis, and averaged values per dendritic segments were pooled. Scanimage software (Pologruto et al., 2003), allowed continuous acquisition of high-magnification images at 0.2-μm intervals at maximal resolution (512 × 512 pixels). Photomultiplier tube voltage settings were maintained at the same level for image collection for each cell. Registered projections of stacks collected were used to determine dendritic spine enumerations. Images of γ-H2AX in CA1 pyramidal neurons were obtained using confocal microscopy (Leica TCS SP8). Images were analyzed by MetaMorph and ImageJ software. All imaging and image analyses were performed in a blind manner.

Statistical analysis

Results are reported as mean ± SEM. The statistical significance was evaluated by one- and two-way ANOVA with post hoc Tukey for multiple comparison. Mann–Whitney U test and Student’s t test were used for two group comparison. Statistical significance was set at p < 0.05 (Table 2).

Table 2.

Statistical table

Graph Data structure Type of test Dataset p value
Fig. 1A Nonparametric Mann-Whitney test AMPAR-EPSC 0.97
NMDAR-EPSC 0.51
AMPAR/NMDAR ratio 0.97
Fig. 1B Nonparametric Mann-Whitney test AMPAR-EPSC 0.69
NMDAR-EPSC 0.47
A/N ratio 0.33
Fig. 1C Nonparametric Mann-Whitney test AMPAR-EPSC 0.79
NMDAR-EPSC 0.79
A/N ratio 0.47
Fig. 1D Parametric Student’s t test PPR of pSup 0.86
PPR of shScr 0.48
PPR of shLuc 0.76
Fig. 2A Nonparametric Mann-Whitney test GABAAR-IPSC 0.21
AMPAR-EPSC 1
AMPAR/GABAAR ratio 0.55
Fig. 2B Nonparametric Mann-Whitney test GABAAR-IPSC 0.43
AMPAR-EPSC 0.41
A/G ratio 0.56
Fig. 2C Nonparametric Mann-Whitney test GABAAR-IPSC 0.69
AMPAR-EPSC 0.92
A/G ratio 0.89
Fig. 2D Nonparametric Mann-Whitney test GABAAR-IPSC 0.15
AMPAR-EPSC 0.24
A/G ratio 0.58
Fig. 2E Parametric Student’s t test PPR of pSup 0.63
PPR of shScr 0.74
PPR of shLuc 0.31
Fig. 3B Parametric One-way ANOVA post hoc Tukey Spine density: pSup vs shLuc 0.45
pSup vs shScr 0.09
shScr vs shLuc 0.65
Spine length: pSup vs shLuc 0.99
pSup vs shScr 0.68
shScr vs shLuc 0.63
Spine width: pSup vs shLuc 0.96
pSup vs shScr 0.37
shScr vs shLuc 0.53
Fig. 4A Parametric Two-way ANOVA post hoc Tukey −100 pA
untrans vs shLuc >0.9999
untrans vs shScr >0.9999
untrans vs pSup >0.9999
shLuc vs shScr >0.9999
shLuc vs pSup >0.9999
shScr vs pSup >0.9999
−50 pA
untrans vs shLuc >0.9999
untrans vs shScr >0.9999
untrans vs pSup >0.9999
shLuc vs shScr >0.9999
shLuc vs pSup >0.9999
shScr vs pSup >0.9999
0 pA
untrans vs shLuc >0.9999
untrans vs shScr >0.9999
untrans vs pSup >0.9999
shLuc vs shScr >0.9999
shLuc vs pSup >0.9999
shScr vs pSup >0.9999
50 pA
untrans vs shLuc >0.9999
untrans vs shScr >0.9999
untrans vs pSup 0.9993
shLuc vs shScr >0.9999
shLuc vs pSup 0.9995
shScr vs pSup 0.9996
100 pA
untrans vs shLuc 0.9997
untrans vs shScr >0.9999
untrans vs pSup 0.9923
shLuc vs shScr 0.9995
shLuc vs pSup 0.9978
shScr vs pSup 0.9928
150 pA
untrans vs shLuc 0.9909
untrans vs shScr >0.9999
untrans vs pSup 0.9964
shLuc vs shScr 0.9959
shLuc vs pSup >0.9999
shScr vs pSup 0.9982
200 pA
untrans vs shLuc 0.9818
untrans vs shScr >0.9999
untrans vs pSup 0.8433
shLuc vs shScr 0.9915
shLuc vs pSup 0.9674
shScr vs pSup 0.8983
250 pA
untrans vs shLuc 0.7984
untrans vs shScr 0.9421
untrans vs pSup 0.4089
shLuc vs shScr 0.9945
shLuc vs pSup 0.8912
shScr vs pSup 0.8019
300 pA
untrans vs shLuc 0.6045
untrans vs shScr 0.8776
untrans vs pSup 0.124
shLuc vs shScr 0.9847
shLuc vs pSup 0.714
shScr vs pSup 0.5448
350 pA
untrans vs shLuc 0.2035
untrans vs shScr 0.8942
untrans vs pSup 0.1894
shLuc vs shScr 0.7588
shLuc vs pSup 0.9865
shScr vs pSup 0.6361
400 pA
untrans vs shLuc 0.1775
untrans vs shScr 0.7432
untrans vs pSup 0.2523
shLuc vs shScr 0.8662
shLuc vs pSup 0.9987
shScr vs pSup 0.844
450 pA
untrans vs shLuc 0.0599
untrans vs shScr 0.8942
untrans vs pSup 0.2917
shLuc vs shScr 0.4825
shLuc vs pSup 0.9928
shScr vs pSup 0.7592
500 pA
untrans vs shLuc 0.0477
untrans vs shScr 0.9921
untrans vs pSup 0.6175
shLuc vs shScr 0.2411
shLuc vs pSup 0.8394
shScr vs pSup 0.8383
550 pA
untrans vs shLuc 0.0101
untrans vs shScr 0.9971
untrans vs pSup 0.8145
shLuc vs shScr 0.0831
shLuc vs pSup 0.4248
shScr vs pSup 0.9297
600 pA
untrans vs shLuc 0.0043
untrans vs shScr >0.9999
untrans vs pSup 0.68
shLuc vs shScr 0.0299
shLuc vs pSup 0.437
shScr vs pSup 0.7694
650 pA
untrans vs shLuc 0.0014
untrans vs shScr 0.995
untrans vs pSup 0.8638
shLuc vs shScr 0.0266
shLuc vs pSup 0.176
shScr vs pSup 0.9617
700 pA
untrans vs shLuc 0.0003
untrans vs shScr 0.9998
untrans vs pSup 0.9417
shLuc vs shScr 0.0061
shLuc vs pSup 0.0571
shScr vs pSup 0.9735
750 pA
untrans vs shLuc 0.0002
untrans vs shScr 0.7664
untrans vs pSup 0.9643
shLuc vs shScr 0.0001
shLuc vs pSup 0.038
shScr vs pSup 0.6549
800 pA
untrans vs shLuc 0.0007
untrans vs shScr 0.5283
untrans vs pSup 0.9959
shLuc vs shScr <0.0001
shLuc vs pSup 0.038
shScr vs pSup 0.6058
850 pA
untrans vs shLuc 0.0041
untrans vs shScr 0.2197
untrans vs pSup >0.9999
shLuc vs shScr <0.0001
shLuc vs pSup 0.0601
shScr vs pSup 0.4478
900 pA
untrans vs shLuc 0.0005
untrans vs shScr 0.4401
untrans vs pSup >0.9999
shLuc vs shScr <0.0001
shLuc vs pSup 0.0174
shScr vs pSup 0.6624
950 pA
untrans vs shLuc 0.0001
untrans vs shScr 0.1954
untrans vs pSup 0.9999
shLuc vs shScr <0.0001
shLuc vs pSup 0.006
shScr vs pSup 0.466
1000 pA
untrans vs shLuc 0.0004
untrans vs shScr 0.1282
untrans vs pSup 0.9992
shLuc vs shScr <0.0001
shLuc vs pSup 0.0108
shScr vs pSup 0.4014
1050 pA
untrans vs shLuc 0.0001
untrans vs shScr 0.2644
untrans vs pSup 0.9762
shLuc vs shScr <0.0001
shLuc vs pSup 0.0025
shScr vs pSup 0.728
1100 pA
untrans vs shLuc 0.0003
untrans vs shScr 0.2803
untrans vs pSup 0.999
shLuc vs shScr <0.0001
shLuc vs pSup 0.0087
shScr vs pSup 0.5944
Fig. 4B Parametric One-way ANOVA post hoc Tukey Threshhold:
shLuc vs shScr 0.0249
shLuc vs pSup 0.0421
shLuc vs untrans 0.0094
shScr vs pSup 0.9958
shScr vs untrans 0.9991
pSup vs untrans 0.9837
Half width:
shLuc vs shScr 0.0002
shLuc vs pSup 0.0165
shLuc vs untrans <0.0001
shScr vs pSup 0.9231
shScr vs untrans 0.9846
pSup vs untrans 0.7637
Fig. 4C Parametric Two-way ANOVA post hoc Tukey -80 mV
shLuc vs shScr >0.9999
shLuc vs pSup >0.9999
shLuc vs untrans >0.9999
shScr vs pSup >0.9999
shScr vs untrans >0.9999
pSup vs untrans >0.9999
-70 mV
shLuc vs shScr >0.9999
shLuc vs pSup 0.9998
shLuc vs untrans >0.9999
shScr vs pSup >0.9999
shScr vs untrans >0.9999
pSup vs untrans >0.9999
-60 mV
shLuc vs shScr >0.9999
shLuc vs pSup 0.9982
shLuc vs untrans 0.9994
shScr vs pSup 0.9995
shScr vs untrans >0.9999
pSup vs untrans 0.9999
−50 mV
shLuc vs shScr 0.9996
shLuc vs pSup 0.9956
shLuc vs untrans 0.998
shScr vs pSup 0.9992
shScr vs untrans 0.9999
pSup vs untrans 0.9999
−40 mV
shLuc vs shScr 0.999
shLuc vs pSup 0.2815
shLuc vs untrans 0.9959
shScr vs pSup 0.3548
shScr vs untrans 0.9999
pSup vs untrans 0.3237
−30 mV
shLuc vs shScr 0.0008
shLuc vs pSup 0.0272
shLuc vs untrans 0.0755
shScr vs pSup 0.7329
shScr vs untrans 0.315
pSup vs untrans 0.9295
−20 mV
shLuc vs shScr 0.0007
shLuc vs pSup 0.0009
shLuc vs untrans 0.0235
shScr vs pSup 0.9999
shScr vs untrans 0.5293
pSup vs untrans 0.5778
−10 mV
shLuc vs shScr 0.0007
shLuc vs pSup 0.0013
shLuc vs untrans 0.0183
shScr vs pSup 0.9983
shScr vs untrans 0.5872
pSup vs untrans 0.7019
0 mV
shLuc vs shScr 0.001
shLuc vs pSup 0.0014
shLuc vs untrans 0.015
shScr vs pSup 0.9998
shScr vs untrans 0.7013
pSup vs untrans 0.7536
10 mV
shLuc vs shScr 0.0005
shLuc vs pSup 0.0008
shLuc vs untrans 0.0083
shScr vs pSup 0.9996
shScr vs untrans 0.6986
pSup vs untrans 0.7654
20 mV
shLuc vs shScr 0.0011
shLuc vs pSup 0.0017
shLuc vs untrans 0.0064
shScr vs pSup 0.9994
shScr vs untrans 0.856
pSup vs untrans 0.9091
30 mV
shLuc vs shScr 0.0015
shLuc vs pSup 0.0017
shLuc vs untrans 0.0042
shScr vs pSup >0.9999
shScr vs untrans 0.9547
pSup vs untrans 0.9613
Fig. 4D Parametric Two-way ANOVA post hoc Tukey −80 mV
shLuc vs shScr >0.9999
shLuc vs pSup >0.9999
shLuc vs untrans >0.9999
shScr vs pSup >0.9999
shScr vs untrans >0.9999
pSup vs untrans >0.9999
−70 mV
shLuc vs shScr >0.9999
shLuc vs pSup 0.9998
shLuc vs untrans >0.9999
shScr vs pSup >0.9999
shScr vs untrans >0.9999
pSup vs untrans >0.9999
−60 mV
shLuc vs shScr >0.9999
shLuc vs pSup 0.9982
shLuc vs untrans 0.9994
shScr vs pSup 0.9995
shScr vs untrans >0.9999
pSup vs untrans 0.9999
−50 mV
shLuc vs shScr 0.9996
shLuc vs pSup 0.9956
shLuc vs untrans 0.998
shScr vs pSup 0.9992
shScr vs untrans 0.9999
pSup vs untrans 0.9999
−40 mV
shLuc vs shScr 0.999
shLuc vs pSup 0.2815
shLuc vs untrans 0.9959
shScr vs pSup 0.3548
shScr vs untrans 0.9999
pSup vs untrans 0.3237
−30 mV
shLuc vs shScr 0.0008
shLuc vs pSup 0.0272
shLuc vs untrans 0.0755
shScr vs pSup 0.7329
shScr vs untrans 0.315
pSup vs untrans 0.9295
−20 mV
shLuc vs shScr 0.0007
shLuc vs pSup 0.0009
shLuc vs untrans 0.0235
shScr vs pSup 0.9999
shScr vs untrans 0.5293
pSup vs untrans 0.5778
−10 mV
shLuc vs shScr 0.0007
shLuc vs pSup 0.0013
shLuc vs untrans 0.0183
shScr vs pSup 0.9983
shScr vs untrans 0.5872
pSup vs untrans 0.7019
0 mV
shLuc vs shScr 0.001
shLuc vs pSup 0.0014
shLuc vs untrans 0.015
shScr vs pSup 0.9998
shScr vs untrans 0.7013
pSup vs untrans 0.7536
10 mV
shLuc vs shScr 0.0005
shLuc vs pSup 0.0008
shLuc vs untrans 0.0083
shScr vs pSup 0.9996
shScr vs untrans 0.6986
pSup vs untrans 0.7654
20 mV
shLuc vs shScr 0.0011
shLuc vs pSup 0.0017
shLuc vs untrans 0.0064
shScr vs pSup 0.9994
shScr vs untrans 0.856
pSup vs untrans 0.9091
30 mV
shLuc vs shScr 0.0015
shLuc vs pSup 0.0017
shLuc vs untrans 0.0042
shScr vs pSup >0.9999
shScr vs untrans 0.9547
pSup vs untrans 0.9613
Fig. 5A Parametric Student’s t test Ctrl vs CPT <0.00001
Fig. 5B Parametric One-way ANOVA post hoc Tukey shLuc vs shScr 0.8937
shLuc vs pSup 0.885
shScr vs pSup 0.9996

Results

Basal excitatory synaptic transmission in shRNA-transfected neurons

To examine the physiologic effects of shRNAs on neuronal function, we prepared organotypic slice cultures from mouse hippocampi and biolistically transfected GFP with pSuper empty vector (pSup), luciferase shRNA (shLuc) and shScr. To test whether control shRNAs display abnormal off-target effect(s) in excitatory synaptic function, simultaneous whole-cell recordings of untransfected and transfected CA1 pyramidal neurons were performed. AMPAR- and NMDAR-mediated evoked EPSCs were measured by stimulating Shaffer collateral inputs (Fig. 1A-C). As expected, transfection of pSuper empty vector, shLuc or shScr exhibited no off-target effects on the amplitude of AMPAR- and NMDAR-EPSCs, as well as the AMPAR/NMDAR ratio, confirming previous results (Hoogenraad et al., 2010; Chen et al., 2014b). Paired-pulse ratio (PPR) measured by the AMPAR-EPSC response obtained by the double stimulation of Shaffer collateral inputs with a 50-ms interval exhibited comparable levels of facilitation in all transfected neurons regardless of the plasmids used, indicating that the transfection of these genes do not affect presynaptic release probability (Fig. 1D).

Figure 1.

Figure 1.

Comparable levels of excitatory synaptic transmission between shRNA-transfected and untransfected neurons. Effect of overexpression of three different plasmid transfections on excitatory synaptic transmission in hippocampal CA1 pyramidal cells. An empty vector (pSup; A), shScr (B), and shLuc (C) were transfected together with pCAG-EGFP. Recordings were conducted 5 d following transfection. Left column, Sample EPSC traces mediated by AMPARs (downward) and NMDARs (upward) from pairs of transfected neurons (Trans) and neighboring untransfected neurons (Untrans). Stimulus artifacts were truncated. Middle columns, Scattered plots of NMDAR- (left) and AMPAR- (right) EPSC amplitude (amp). Each pair of transfected and neighboring untransfected cells are presented as open symbols while filled symbols indicate the mean. Right column, Bar graphs of AMPAR/NMDAR ratios. D, PPR of AMPAR-EPSCs recorded from trans- and untransfected neurons, as indicated. Left, Sample traces. Normalized EPSCs to the first EPSC amplitude from trans- (gray) and untransfected neurons (black) are superimposed. Right, Summary graphs of PPR. The PPR was calculated by dividing the average amplitude of the second EPSC by that of the first EPSC. Number of cell pairs tested: pSup, 10 cells/6 mice; shLuc, 10/6; shScr, 11/6 (for AMPAR-EPSC, NMDAR-EPSC, and PPR, respectively).

Normal inhibitory synaptic transmission in shRNA-transfected neurons

Next, we examined the effects of shRNAs on inhibitory synaptic transmission and the balance of excitatory and inhibitory transmission. Neither GABAAR-mediated IPSC nor the ratio of AMPAR and GABAAR responses displayed significant changes following transfection of any of the three plasmids (Fig. 2A–C). Paired-pulse stimulation of inhibitory inputs exhibited comparable levels of depression in all transfected neurons, suggesting that the transfection of these genes do not have any effects on presynaptic release probability (Fig. 2E).

Figure 2.

Figure 2.

Comparable level of inhibitory synaptic transmission between shRNA-transfected and untransfected neurons. Effect of overexpression of the three different plasmid on excitatory and inhibitory synaptic transmission in hippocampal CA1 pyramidal cells. An empty vector (pSup; A), shScr (B), shLuc (C) and shLuc x3 (triple the amount of shLuc) (D) were transfected together with pCAG-EGFP. Left column, Sample AMPAR-EPSC and GABAAR-IPSC traces mediated by AMPARs (downward) and GABAARs (upward) from pairs of transfected neurons (Trans) and neighboring untransfected neurons (Untrans). Stimulus artifacts were truncated. Middle columns, Scattered plots of GABAAR- (left) and AMPAR- (right) EPSC amplitude (amp). Each pair of transfected and neighboring untransfected cells are presented as open symbols while filled symbols indicate the mean. Right column, Bar graphs of AMPAR/GABAAR ratios. E, PPR of GABAAR-IPSCs recorded from trans- and untransfected neurons, as indicated. Left, Sample traces. Normalized IPSCs to the first IPSC amplitude from trans- (gray) and untransfected neurons (black) were superimposed. The First GABAAR-IPSC overlaps with the second IPSC. Therefore, the first EPSC was cancelled by subtracting the traces receiving a single pulse from those receiving a paired pulse, both normalized to the first response. Right, Summary graphs of PPR. The PPR was calculated by dividing the average amplitude of the second IPSC by that of the first IPSC. Number of cell pairs: pSup, 13, 11, and 13 cells/7 mice (13, 11, 13/7); shLuc (18, 13, 18/8); shScr (20, 15, 20/8); shLuc x3 (13, 12/6), for GABAAR-IPSC, AMPAR-EPSC, and PPR, respectively.

In contrast to our results, Alvarez et al. (2006) reported that shLuc displayed reduced inhibitory and excitatory synaptic transmission in organotypic hippocampal CA1 neurons. The dosage of shRNAs, reported to contribute to off-target effects (Jackson et al., 2006a), may account for this discrepancy as Alvarez et al. (2006) used a 1.3-fold larger amount of shRNA than used in the present study. To test this possibility, we increased the shLuc plasmid concentration three-fold and measured IPSCs and AMPAR-EPSCs (Fig. 2D). A three-fold increase of shLuc did not significantly reduce synaptic transmission, indicating that a potential dosage-dependent off-target effect of shLuc is small in synaptic function as measured here and most likely is not the primary reason for the differences in results between those reported here and by Alvarez et al. (2006).

Normal dendritic morphology in shRNA-transfected neurons

We next addressed the effects of shRNAs on neuronal morphology. We overexpressed shRNAs together with EGFP. The transfected neurons were immunostained against GFP followed by imaging using two-photon microscopy (Fig. 3). The dendritic morphology, including spine density and dimension, was found to be unchanged by overexpression of the two shRNAs. Overall, we did not identify off-target effects of control shRNAs in synaptic function and structure.

Figure 3.

Figure 3.

Comparable level of dendritic structure between different shRNA-transfected neurons. Effect of overexpression of three different plasmid transfections on dendritic morphology in hippocampal CA1 pyramidal neurons. A, Low (top) and high (bottom) magnification images obtained from empty vector, pSup (left), shLuc (middle), and shScr (right) transfected neurons. Each plasmid was cotransfected with pCAG-EGFP. Fixed slices were immunostained against GFP and neuronal images were obtained by two-photon microscopy. B, Scatter plots of spine length (left), width (middle), and density (right). Error bars indicate SEMs. Note that none of these parameters displayed statistical significance by gene transfection. Number of dendritic segments/cells/mice: pSup, 31/5/5; shLuc, 31/5/5; and shScr, 28/5/5.

Reduced membrane excitability and voltage-gated ion channel function in shLuc-transfected CA1 pyramidal neurons

To examine the effects of shRNAs on membrane properties, neurons transfected with shRNAs and EGFP were compared to untransfected neurons with respect to action potentials (APs) and basic membrane properties (Fig. 4; Table 1). To our great surprise, CA1 neurons transfected with shLuc exhibited a reduced number of APs compared to untransfected control neurons (Fig. 4A). The transfection of pSup did not cause any abnormal excitability, indicating that this vector backbone is not the cause of the reduced excitability in shLuc-transfected neurons.

Figure 4.

Figure 4.

Reduced membrane excitabilities in shLuc-transfected CA1 pyramidal neurons. A, Effect of shRNA overexpression on neuronal excitability. An empty vector (pSup), shLuc, or shScr was transfected together with pCAG-EGFP. Top, Sample traces from untransfected and transfected CA1 pyramidal neurons in organotypic hippocampal slice cultures. The superimposed traces were elicited by current injections of 0, 100, and 500 pA for 200 ms. Bottom, Summary graph of the frequency of APs in untransfected and transfected neurons. The input–output relationship [number of spikes elicited versus amount of current injection (200-ms duration)] was plotted for untransfected and transfected neurons. Neurons were held at resting membrane potentials (−56.8 mV ± 0.57, n = 100 cells, 9 mice). Number of cells tested: untrans, 45 cells from eight mice (45/8); pSup, 13/7; shScr, 20/5; and shLuc, 22/9. B, Effect of shRNA overexpression on AP kinetics. Top, Sample traces from CA1 pyramidal neurons untransfected and transfected with the three different plasmids. Single APs were induced by current injection (100 pA for 4 ms) and threshold (horizontal arrows) and half width of AP determined by vertical double arrow heads in trans- and untransfected neurons were measured. Bottom, Summary graph of the threshold (left) and half width (right) of single AP in untransfected and transfected neurons. Number of cells tested: untrans, 45 cells/8 mice (45/8); pSup, 13/7; shScr, 20/5; shLuc, 22/9. C, top, Sample traces of sodium currents recorded from untransfected and transfected CA1 pyramidal neurons. Note that these currents were completely blocked by TTX (right). Bottom, Summary graph of the sodium currents in untransfected and transfected neurons. Neurons were voltage-clamped at –80 mV and depolarized from –80 to 30 mV (10-ms duration). Number of cells: untrans, 16 cells/5 mice; pSup, 10/5; shScr, 10/5; shLuc, 14/6. D, top, Sample traces of potassium currents recorded from untransfected and transfected CA1 pyramidal neurons. Note that these upward currents were completely blocked by TEA and 4-AP (right). Bottom, Summary graph of the potassium currents in untransfected and transfected neurons. Neurons were voltage-clamped at –80 mV and depolarized from –80 to 30 mV (10-ms duration). Number of cells tested: untrans, 15 cells/4 mice; pSup, 13/3; shScr, 7/3; shLuc, 9/3.

Table 1.

The basic membrane properties of untransfected and gene transfected hippocampal CA1 neurons

Cell types Peak amplitude of AP (mV) Resting membrane potential (mV) Input resistance (MΩ) Series resistance (MΩ) Number of cells/mice
untrans 106.1 ± 1.331 −55.87 ± 0.900 105.6 ± 8.452 8.127 ± 0.2768 45/8
pSup 108.5 ± 2.452 −58.51 ± 1.604 122.9 ± 18.62 8.869 ± 0.6093 13/7
shScr 109.3 ± 1.456 −57.47 ± 1.104 143.1 ± 18.03 8.456 ± 0.3902 20/5
shLuc 106.0 ± 2.224 −57.07 ± 1.054 131.0 ± 11.29 9.349 ± 0.3110 22/9
Statistics p value shLuc vs shScr 0.7606 shLuc vs shScr >0.9999 shLuc vs shScr 0.9893 shLuc vs shScr 0.4875
shLuc vs pSup 0.9657 shLuc vs pSup 0.9777 shLuc vs pSup 0.9996 shLuc vs pSup 0.9755
shLuc vs untrans >0.9999 shLuc vs untrans 0.9604 shLuc vs untrans 0.567 shLuc vs untrans 0.0681
shScr vs untrans 0.6744 shScr vs pSup 0.9958 shScr vs pSup 0.9434 shScr vs pSup 0.988
shScr vs pSup >0.9999 shScr vs untrans 0.8504 shScr vs untrans 0.1487 shScr vs untrans 0.9813
pSup vs untrans 0.9603 pSup vs untrans 0.5991 pSup vs untrans 0.9558 pSup vs untrans 0.7507

Note that all of these parameters were not statistically different between four cell groups. Statistics were done by one-way ANOVA with post hoc Tukey.

To further test whether shLuc changed the threshold and kinetics of APs, we compared the onset, threshold, amplitude and half-width of APs in untransfected and transfected neurons (Fig. 4B; Table 1). We found that APs in shLuc-transfected neurons exhibited increased half-width (Fig. 4B, lower right) and decrease of threshold (Fig. 4B, lower left) without changing the amplitude of AP (Table 1). The basic membrane properties, including resting membrane potential and series and input resistance, remained unchanged, indicating an off-target effect of shLuc on voltage-gated ion channel function (Table 1). Taken together, our results strongly suggest that shLuc has off-target effect(s) on membrane excitability and AP kinetics in CA1 pyramidal neurons.

Given these striking results, we next addressed the underlying mechanism of the off-target effect of shLuc on membrane excitability and APs. Whole-cell voltage clamp recordings were performed in CA1 pyramidal neurons and the current−voltage (I-V) relationship of TTX-sensitive sodium (Na+) channel current was measured. Importantly, Na+ current maximum amplitude between −20 and +30 mV was markedly reduced in shLuc-transfected compared to untransfected neurons (Fig. 4C, closed circles). We then tested whether voltage-dependent potassium (K+) currents were altered by shLuc transfection. The potassium channel currents, sensitive to the broad potassium channel blockers, TEA, and 4-AP, were also markedly reduced in shLuc-transfected neurons compared with other neurons (Fig. 4D, closed circle). These results indicate that shLuc has off-target effects on voltage-gated sodium and potassium channels.

Normal H2A-X phosphorylation in shRNA-transfected neurons

Lastly, we addressed whether the reduced membrane excitability in shLuc-transfected neurons is due to potential cellular stress or damage to cellular functions. To test this hypothesis, we examined the level of the phosphorylation of histone H2A.X. H2A.X phosphorylation at serine 139 (γ-H2AX) is a routinely used biomarker to detect DNA damage and DNA replication stress (Mah et al., 2010). γ-H2AX is greatly increased in neurons by seizure insult (Crowe et al., 2011) and camptothecin, an inhibitor of the DNA enzyme topoisomerase I (Mah et al., 2010; Fig. 5A). Importantly, the level of γ-H2AX in shLuc-transfected neurons was comparable to that of pSup- and shScr-transfected neurons, suggesting that transfection of shLuc does not alter the overall health of the neurons (Fig. 5B). In summary, we conclude that shLuc exhibits off-target effects on membrane excitability without changing synaptic transmission, dendritic morphology or cellular health.

Figure 5.

Figure 5.

Comparable levels of H2A-X phosphorylation in neurons transfected with different shRNA vectors. γ-H2AX immunoreactivity in shRNA-transfected hippocampal CA1 neurons. A, left, Representative confocal images of immunofluorescence staining against γ-H2AX (red) and DAPI (blue) in CA1 pyramidal neurons. Right, Quantification of γ-H2AX signal intensity in nuclei. Treatment with 10 μM camptothecin (CPT) for 6 h increased the signal of γ-H2AX in nuclei of neurons in organotypic slice cultures, confirming the specificity of the anti-γ-H2AX antibody. AU, arbitrary units. Number of cells/mice: mock control (Ctrl), 140/3; CPT, 133/3. B, left, Representative confocal images of immunofluorescence staining against γ-H2AX (red), GFP epifluorescence (green) and DAPI (blue) in pSup-, shLuc-, and shScr-transfected neurons. Right, Quantification of γ-H2AX signal intensity in nuclei. All transfected neurons exhibited comparable levels of γ-H2AX signal. Number of cells/mice: pSup, 11/3; shScr, 7/4; shLuc, 7/3. Scale bars: 10 μm.

Discussion

RNAi is a powerful approach to validate the function of target genes both in vitro and in vivo. In light of repeated indications that shRNAs may have off-target effects (Cullen, 2006), their effects should be carefully studied. Here, we found that shLuc transfection increased AP threshold and half-width in addition to reducing the number of APs. shLuc also reduced TTX-sensitive sodium and TEA- and 4AP-sensitive potassium currents that contribute to shaping APs. In contrast to its effects on these currents, shLuc failed to alter basic membrane properties, suggesting that shLuc does not exhibit off-target effects on pumps (e.g., Na+/K+-ATPase) and other channels (e.g., leak channels) that contribute to maintaining resting membrane potential.

Is there any possibilities that shLuc targets the genes regulating the function of voltage-gated ion channels? A BLAST search against shLuc guide sequence (19 base-pairs: TCGAAGTATTCCGCGTACG), the strand incorporated into RNA-induced silencing complex, displayed seven genes with low homology (11–12 bp match, 58–63% homology), including Dnah11 (dynein axonemal heavy chain 11, XM_017314949.1), Plpp5 (phospholipid phosphatase 5, NM_001293703.1), Lpcat2 (lysophosphatidylcholine acyltransferase 2, XM_006531052.2), Asb7 (Ankyrin repeat and SOCS box-containing 7, XM_006540568.3), Ipo8 (importin 8, XR_001785142.1), Dis3 (DIS3 homolog exosome endoribonuclease and 3´->5´ exoribonuclease, XM_006519602.3), Fbcl20 (F-box and leucine-rich repeat protein 20, XM_006534299.3). Based on the homology, it is difficult to consider these genes as the potential target of shLuc and none of these have been known as the regulators of Na+ and K+ channels.

Alvarez et al. (2006) reported that shLuc caused a lack of dendritic spines and reduced excitatory and inhibitory synaptic transmission in hippocampal pyramidal neurons. However, we were unable to confirm their observations. As described in our results, neurons transfected with shLuc display normal synaptic transmission and dendritic spine structure compared with untransfected and other genes transfected neurons. Our results are also supported by other studies demonstrating normal excitatory synaptic transmission, long-term potentiation and dendritic morphology in shLuc-transfected neurons (Hoogenraad et al., 2010; Wakita et al., 2011; Chen et al., 2014a,b). Since Alvarez et al. (2006) used the same experimental approach as ours (Stoppini et al., 1991), this discrepancy may be due to differences in precise details of the experimental procedures, including the vector backbone and the methods of gene transfection in organotypic slice cultures. For example, Alvarez et al. (2006) used U6 promoter to synthesize shLuc but pSuper vector consists with H1 promoter. Both H1 and U6 promoters are member of the type III class of Pol III RNA promoters, but it is possible that these two different promoters cause the distinct off-target effects in neurons (Mäkinen et al., 2006). However, regardless of this diverging result, it is clear that shLuc exhibits off-target effects on ion channel or synaptic function, discouraging the usage of shLuc as a control shRNA, at least in hippocampal CA1 pyramidal cells.

While comprehensive guidelines for the RNAi approach have been described (2003), detailed studies of off-target effects of control shRNAs are very limited. Thus, our comparative study on shRNAs in pyramidal neurons is an important consideration for future RNAi experiments. Of note, gene expression is differentially regulated by cell-type, cellular activity and development. Therefore, careful evaluation is required for shRNA constructs, including nonsilencing controls. However, we consider that shScr is currently the best validated control for the RNAi approach in mammalian cells including neurons.

Acknowledgments

Acknowledgements: We thank Dr. Paul D. Gardner for valuable discussions and Mrs. Naoe Watanabe for skillful technical assistance.

Synthesis

Reviewing Editor: Christophe Bernard, INSERM & Institut de Neurosciences des Systèmes

Decisions are customarily a result of the Reviewing Editor and the peer reviewers coming together and discussing their recommendations until a consensus is reached. When revisions are invited, a fact-based synthesis statement explaining their decision and outlining what is needed to prepare a revision will be listed below. The following reviewer(s) agreed to reveal their identity: Weinan Sun.

Both reviewers noted the high quality of the data presented here.

The result is very important for the community, and we agree that at this stage, it may not be necessary to provide a mechanism. However, it is essential for this kind of study not to add confusion to the field.

1. We agree that it is important for you to test if the cells transfected with shLuc are as healthy as controls. Knowing that shLuc damages or not the cells would be really useful.

2. The lack of consistency with Alvarez et al., 2006 raises important issues. Since you propose that dosage may be an explanation, would it be possible to perform a dose-response curve?

Minor

Methods section:

“35 C{degree sign}” should be “35 {degree sign}C”.

post hoc “Turkey” should be “Tukey”.

The n value for statistical analysis of Figure 3 needs to be specified. Authors refer to 5 cells per 5 mice per group, however the number of dots represented in the graphs seems a lot larger.

Figure 4D should indicate TEA not TAE.

The statistical analysis described in table 1 should be included in the statistics table with p values.

Table 1 should read shLuc not shLuci.

References

  1. (2003) Whither RNAi? Nat Cell Biol 5:489–490. [DOI] [PubMed] [Google Scholar]
  2. Alvarez VA, Ridenour DA, Sabatini BL (2006) Retraction of synapses and dendritic spines induced by off-target effects of RNA interference. J Neurosci 26:7820–7825. 10.1523/JNEUROSCI.1957-06.2006 [DOI] [PMC free article] [PubMed] [Google Scholar]
  3. Birmingham A, Anderson EM, Reynolds A, Ilsley-Tyree D, Leake D, Fedorov Y, Baskerville S, Maksimova E, Robinson K, Karpilow J, Marshall WS, Khvorova A (2006) 3' UTR seed matches, but not overall identity, are associated with RNAi off-targets. Nat Methods 3:199–204. 10.1038/nmeth854 [DOI] [PubMed] [Google Scholar]
  4. Chen Y, Wang Y, Modrusan Z, Sheng M, Kaminker JS (2014a) Regulation of neuronal gene expression and survival by basal NMDA receptor activity: a role for histone deacetylase 4. J Neurosci 34:15327–15339. [DOI] [PMC free article] [PubMed] [Google Scholar]
  5. Chen Y, Wang Y, Erturk A, Kallop D, Jiang Z, Weimer RM, Kaminker J, Sheng M (2014b) Activity-induced Nr4a1 regulates spine density and distribution pattern of excitatory synapses in pyramidal neurons. Neuron 83:431–443. [DOI] [PubMed] [Google Scholar]
  6. Crowe SL, Tsukerman S, Gale K, Jorgensen TJ, Kondratyev AD (2011) Phosphorylation of histone H2A.X as an early marker of neuronal endangerment following seizures in the adult rat brain. J Neurosci 31:7648–7656. 10.1523/JNEUROSCI.0092-11.2011 [DOI] [PMC free article] [PubMed] [Google Scholar]
  7. Cullen BR (2006) Enhancing and confirming the specificity of RNAi experiments. Nat Methods 3:677–681. 10.1038/nmeth913 [DOI] [PubMed] [Google Scholar]
  8. Fire A, Xu S, Montgomery MK, Kostas SA, Driver SE, Mello CC (1998) Potent and specific genetic interference by double-stranded RNA in Caenorhabditis elegans . Nature 391:806–811. 10.1038/35888 [DOI] [PubMed] [Google Scholar]
  9. Futai K, Doty CD, Baek B, Ryu J, Sheng M (2013) Specific trans-synaptic interaction with inhibitory interneuronal neurexin underlies differential ability of neuroligins to induce functional inhibitory synapses. J Neurosci 33:3612–3623. 10.1523/JNEUROSCI.1811-12.2013 [DOI] [PMC free article] [PubMed] [Google Scholar]
  10. Futai K, Kim MJ, Hashikawa T, Scheiffele P, Sheng M, Hayashi Y (2007) Retrograde modulation of presynaptic release probability through signaling mediated by PSD-95-neuroligin. Nat Neurosci 10:186–195. 10.1038/nn1837 [DOI] [PMC free article] [PubMed] [Google Scholar]
  11. Grimm D, Streetz KL, Jopling CL, Storm TA, Pandey K, Davis CR, Marion P, Salazar F, Kay MA (2006) Fatality in mice due to oversaturation of cellular microRNA/short hairpin RNA pathways. Nature 441:537–541. 10.1038/nature04791 [DOI] [PubMed] [Google Scholar]
  12. Hoogenraad CC, Popa I, Futai K, Sanchez-Martinez E, Wulf PS, van Vlijmen T, Dortland BR, Oorschot V, Govers R, Monti M, Heck AJ, Sheng M, Klumperman J, Rehmann H, Jaarsma D, Kapitein LC, van der Sluijs P (2010) Neuron specific Rab4 effector GRASP-1 coordinates membrane specialization and maturation of recycling endosomes. PLoS Biol 8:e1000283. 10.1371/journal.pbio.1000283 [DOI] [PMC free article] [PubMed] [Google Scholar]
  13. Jackson AL, Bartz SR, Schelter J, Kobayashi SV, Burchard J, Mao M, Li B, Cavet G, Linsley PS (2003) Expression profiling reveals off-target gene regulation by RNAi. Nat Biotechnol 21:635–637. 10.1038/nbt831 [DOI] [PubMed] [Google Scholar]
  14. Jackson AL, Burchard J, Schelter J, Chau BN, Cleary M, Lim L, Linsley PS (2006a) Widespread siRNA "off-target" transcript silencing mediated by seed region sequence complementarity. RNA 12:1179–1187. [DOI] [PMC free article] [PubMed] [Google Scholar]
  15. Jackson AL, Burchard J, Leake D, Reynolds A, Schelter J, Guo J, Johnson JM, Lim L, Karpilow J, Nichols K, Marshall W, Khvorova A, Linsley PS (2006b) Position-specific chemical modification of siRNAs reduces "off-target" transcript silencing. RNA 12:1197–1205. [DOI] [PMC free article] [PubMed] [Google Scholar]
  16. Kariko K, Bhuyan P, Capodici J, Weissman D (2004) Small interfering RNAs mediate sequence-independent gene suppression and induce immune activation by signaling through toll-like receptor 3. J Immunol 172:6545–6549. 10.4049/jimmunol.172.11.6545 [DOI] [PubMed] [Google Scholar]
  17. Khan AA, Betel D, Miller ML, Sander C, Leslie CS, Marks DS (2009) Transfection of small RNAs globally perturbs gene regulation by endogenous microRNAs. Nat Biotechnol 27:549–555. 10.1038/nbt.1543 [DOI] [PMC free article] [PubMed] [Google Scholar]
  18. Lo DC, McAllister AK, Katz LC (1994) Neuronal transfection in brain slices using particle-mediated gene transfer. Neuron 13:1263–1268. [DOI] [PubMed] [Google Scholar]
  19. Mah LJ, El-Osta A, Karagiannis TC (2010) gammaH2AX: a sensitive molecular marker of DNA damage and repair. Leukemia 24:679–686. 10.1038/leu.2010.6 [DOI] [PubMed] [Google Scholar]
  20. Mäkinen PI, Koponen JK, Kärkkäinen AM, Malm TM, Pulkkinen KH, Koistinaho J, Turunen MP, Ylä-Herttuala S (2006) Stable RNA interference: comparison of U6 and H1 promoters in endothelial cells and in mouse brain. J Gene Med 8:433–441. 10.1002/jgm.860 [DOI] [PubMed] [Google Scholar]
  21. McAllister AK (2000) Biolistic transfection of neurons. Sci STKE 2000:pl1. 10.1126/stke.2000.51.pl1 [DOI] [PubMed] [Google Scholar]
  22. Pologruto TA, Sabatini BL, Svoboda K (2003) ScanImage: flexible software for operating laser scanning microscopes. Biomed Eng Online 2:13. 10.1186/1475-925X-2-13 [DOI] [PMC free article] [PubMed] [Google Scholar]
  23. Sledz CA, Holko M, de Veer MJ, Silverman RH, Williams BR (2003) Activation of the interferon system by short-interfering RNAs. Nat Cell Biol 5:834–839. 10.1038/ncb1038 [DOI] [PubMed] [Google Scholar]
  24. Stoppini L, Buchs PA, Muller D (1991) A simple method for organotypic cultures of nervous tissue. J Neurosci Methods 37:173–182. [DOI] [PubMed] [Google Scholar]
  25. Wakita Y, Kakimoto T, Katoh H, Negishi M (2011) The F-BAR protein Rapostlin regulates dendritic spine formation in hippocampal neurons. J Biol Chem 286:32672–32683. 10.1074/jbc.M111.236265 [DOI] [PMC free article] [PubMed] [Google Scholar]
  26. Zhang H, Macara IG (2006) The polarity protein PAR-3 and TIAM1 cooperate in dendritic spine morphogenesis. Nat Cell Biol 8:227–237. 10.1038/ncb1368 [DOI] [PubMed] [Google Scholar]

Articles from eNeuro are provided here courtesy of Society for Neuroscience

RESOURCES