Summary

Acute Mouse Brain Slicing to Investigate Spontaneous Hippocampal Network Activity

Published: August 28, 2020
doi:

Summary

This protocol describes the preparation of horizontal hippocampal-entorhinal cortex (HEC) slices from mice exhibiting spontaneous sharp-wave ripple activity. Slices are incubated in a simplified interface holding chamber and recordings are performed under submerged conditions with fast-flowing artificial cerebrospinal fluid to promote tissue oxygenation and the spontaneous emergence of network-level activity.

Abstract

Acute rodent brain slicing offers a tractable experimental approach to gain insight into the organization and function of neural circuits with single-cell resolution using electrophysiology, microscopy, and pharmacology. However, a major consideration in the design of in vitro experiments is the extent to which different slice preparations recapitulate naturalistic patterns of neural activity as observed in vivo. In the intact brain, the hippocampal network generates highly synchronized population activity reflective of the behavioral state of the animal, as exemplified by the sharp-wave ripple complexes (SWRs) that occur during waking consummatory states or non-REM sleep. SWRs and other forms of network activity can emerge spontaneously in isolated hippocampal slices under appropriate conditions. In order to apply the powerful brain slice toolkit to the investigation of hippocampal network activity, it is necessary to utilize an approach that optimizes tissue health and the preservation of functional connectivity within the hippocampal network. Mice are transcardially perfused with cold sucrose-based artificial cerebrospinal fluid. Horizontal slices containing the hippocampus are cut at a thickness of 450 μm to preserve synaptic connectivity. Slices recover in an interface-style chamber and are transferred to a submerged chamber for recordings. The recording chamber is designed for dual surface superfusion of artificial cerebrospinal fluid at a high flow rate to improve oxygenation of the slice. This protocol yields healthy tissue suitable for the investigation of complex and spontaneous network activity in vitro.

Introduction

Electrophysiological measurement from living hippocampal slices in vitro is a powerful experimental approach with numerous advantages. The experimenter can use a microscope, micromanipulators, and a recording system to directly visualize and collect measurements from individual neurons in the tissue. Tissue slices are also very accessible to photostimulation or drug delivery for optogenetic, chemogenetic, or pharmacological experiments.

The hippocampal network generates highly synchronous population activity in vivo, visible as oscillations in the extracellular local field potential1,2,3,4,5. Brain slice methods have been leveraged to gain insight into the cellular and circuit mechanisms underlying these neuronal network oscillations. Foundational work from Maier et al. demonstrated that sharp wave-ripple complexes (SWRs) can emerge spontaneously in slices of the ventral hippocampus6,7. Subsequent studies from multiple investigators have gradually elucidated many aspects of SWRs, including the role of neuromodulators in regulating the network state of the hippocampus8,9,10 and the synaptic mechanisms that drive the in vitro reactivation of neuronal ensembles previously active during behavior in vivo11. Brain slice experiments have also provided insight into the gamma range oscillation (30–100 Hz), a distinct hippocampal network state believed to support memory encoding and recall12,13. Finally, recognizing the central role of the hippocampus and associated structures in the pathophysiology of temporal lobe epilepsy14,15, researchers have used hippocampal slice preparations to investigate the generation and propagation of epileptiform activity. Carter et al. demonstrated that combined hippocampal-entorhinal cortex slices prepared from chronically epileptic animals can spontaneously generate epileptiform discharges in vitro16. Subsequently, Karlócai et al. explored the mechanisms underlying epileptiform discharges in hippocampal slices by using modified artificial cerebrospinal fluid (ACSF) with altered ion concentrations (reduced Mg2+ or elevated K+) or added drugs (4AP or gabazine)17.

Investigators have developed numerous hippocampal slice approaches that differ in key ways: (1) the region of the hippocampus contained in the slice (dorsal, intermediate, or ventral); (2) the presence or absence of extrahippocampal tissues such as the entorhinal cortex; (3) the orientation used to cut slices (coronal, sagittal, horizontal, or oblique); and (4) the conditions under which the tissue is maintained after slicing (submerged fully in ACSF or held at the interface of ACSF and humidified, carbogen-rich air).

The choice of which slicing approach to use should be determined by the experimental objective. For example, transverse or coronal slices of the dorsal hippocampus maintained under submerged conditions have been used very effectively for the investigation of intrahippocampal circuitry and synaptic plasticity18,19,20. However, such preparations do not spontaneously generate network oscillations as readily as slices from the ventral hippocampus21,22,23. Although a state of persistent SWR activity can be induced by tetanic stimulation in transverse slices from the dorsal and ventral hippocampus24, spontaneous SWRs are more readily observed in ventral slices7,25.

An inherent physiological and anatomical distinction between the dorsal and ventral hippocampus is supported by studies performed both in vivo and in vitro26. Recordings in rats revealed strongly coherent theta rhythms throughout the dorsal and intermediate hippocampus, yet poor coherence between the ventral region and the rest of the hippocampus27. SWRs in vivo propagate readily between the dorsal and intermediate hippocampus, while SWRs that originate in the ventral hippocampus often remain local28. The associational projections originating from CA3 pyramidal neurons that reside in the dorsal and intermediate hippocampus project long distances along the longitudinal axis of the hippocampus. CA3 projections originating from ventral regions remain relatively local, and thus are less likely to be severed during the slicing process29,30. Ventral slices may, therefore, better preserve the recurrent network necessary to generate population synchrony. The propensity of ventral slices to generate spontaneous network activities in vitro may also reflect higher intrinsic excitability of pyramidal neurons or weaker GABAergic inhibition in the ventral hippocampus as compared to more dorsal regions31. Indeed, ventral hippocampal slices are more susceptible to epileptiform activity32,33. Thus, many studies of spontaneous physiological8,9,11,24 or pathological16,34,35,36 network oscillations have traditionally used a horizontal slicing approach, sometimes with a slight angle in the fronto-occipital direction, which yields tissue slices parallel to the transverse plane of the ventral hippocampus.

Network connectivity is unavoidably impacted by the slicing procedure as many cells in the slice will be severed. The angle and thickness of the slice and the tissue retained in the preparation should be considered to optimize connectivity in the circuits of interest. Many studies have utilized horizontal combined hippocampal-entorhinal cortex slices (HEC) to explore interactions between the two structures in the context of physiological or pathological network oscillations. Roth et al. performed dual recordings from the CA1 subfield of the hippocampus and layer V of the medial entorhinal cortex to demonstrate propagation of SWR activity through the HEC slice37. Many studies of epileptiform activity have used the HEC slice preparation to investigate how epileptiform discharges propagate through the corticohippocampal network16,35,36,38. It is important to note that preservation of the intact corticohippocampal loop is not a prerequisite for spontaneous SWRs, epileptiform discharges, or gamma oscillations; network oscillations can be generated in transverse slices of the dorsal or ventral hippocampus with no attached parahippocampal tissues21,22,23, 25,39,40,41. A more important factor for the spontaneous generation of network oscillations in hippocampal slices may be the thickness of each slice, as a thicker slice (400–550 μm) will preserve more connectivity in the CA2/CA3 recurrent network21,22,25.

Although angled horizontal HEC slices (cut with an approximately 12° angle in the fronto-occipital direction) have been used to study the functional connectivity of the corticohippocampal loop11,16,34,35,42, such angled preparations are not required for spontaneous network activity43,44,45. However, the use of an angled slicing plane does allow the investigator to selectively make slices that best preserve the transversely-oriented lamellae of either the ventral or intermediate hippocampus, depending on whether a downward or an upward angle is applied (Figure 1). This approach is conceptually similar to that used by Papatheodoropoulos et al., 2002, who dissected each hippocampus free and then used a tissue chopper to create transverse slices along the entire dorsal-ventral axis21. In the light of the aforementioned functional distinctions between the ventral and dorsal-intermediate hippocampus, investigators should consider the anatomical origin of slices when designing experiments or interpreting results. Using an agar ramp during the slicing procedure is a simple way to preferentially produce slices from either the intermediate or ventral hippocampus.

Hippocampal slices can be maintained in either a submerged chamber (with the tissue fully immersed in ACSF), or an interface-style chamber (e.g., Oslo or Haas chamber, with slices covered only by a thin film of flowing media). Interface maintenance enhances oxygenation of the tissue, which promotes neuronal survival and allows for sustained high levels of interneuronal activity. Traditionally, submerged recording conditions use a slower ACSF flow rate that does not provide adequate tissue oxygenation for stable expression of network-level oscillations. In submerged hippocampal slices carbachol-induced gamma oscillations are only observed transiently46,47, while they can be stably maintained in interface recording chambers10,48,49. As such, many studies of complex spontaneous activity in vitro have relied on interface recording chambers to investigate sharp-wave ripple complexes6,7,8,9,10,25,37, gamma oscillations10,13, and epileptiform activity16,38,45,47.

In a submerged-style recording chamber, an immersion microscope objective can be used to visualize individual cells and selectively target healthy-looking cells for recordings. The submerged preparation also allows fine control over the cellular milieu, as submersion facilitates rapid diffusion of drugs or other compounds to the tissue. Thus, a modified methodology in which stable network oscillations are maintained under submerged conditions represents a powerful experimental approach. This approach is exemplified by the work of Hájos et al., in which hippocampal slices recover in a simplified interface-style holding chamber for several hours before transfer to a modified submerged recording chamber with a high flow rate of ACSF (~6 mL/min) to enhance oxygen supply to the tissue12,48,49. Under these conditions, high levels of interneuron activity and stable spontaneous network oscillations can be maintained in a submerged recording chamber. This modified approach allows the investigators to perform visually guided whole-cell patch clamp recordings and characterize the contribution of morphologically identified cell types to carbachol-induced gamma oscillations12. SWRs can also occur spontaneously in submerged hippocampal slices with a fast flow rate of ACSF11,48,49. Maier et al. demonstrated that hippocampal slices that recovered in an interface chamber before transfer to a submerged recording chamber reliably exhibited spontaneous SWRs, whereas slices that recovered submerged in a beaker before transfer to a submerged recording chamber showed smaller evoked field responses, lower levels of spontaneous synaptic currents, and only very rarely exhibited spontaneous SWRs43. Schlingloff et al. used this improved methodology to demonstrate the role of parvalbumin-expressing basket cells in the generation of spontaneous SWRs44.

The following protocol presents a slicing method through which spontaneously active neurons in horizontal hippocampal slices can be recovered under interface conditions and subsequently maintained in a submerged recording chamber suitable for pharmacological or optogenetic manipulations and visually guided recordings.

Protocol

All methods described here have been approved by the Institutional Animal Care and Use Committee at Columbia University (AC-AAAU9451). 1. Prepare solutions Prepare sucrose cutting solution for slicing as described in Table 1. NOTE: After preparing 1 L of sucrose solution, freeze a small amount (approximately 100–200 mL) in an ice tray. These frozen sucrose ice cubes will be blended into an icy slurry (see step 4.3). Prepare a…

Representative Results

Presented here are representative recordings from HEC slices prepared as described in this protocol. Following recovery in an interface holding chamber (Figure 1C), slices are transferred individually to a submerged recording chamber (Figure 2B). The recording chamber is supplied with carbogen-saturated ACSF using a peristaltic pump (Figure 2A). The pump first draws ACSF from a holding beaker into a heated reservoir. Carbogen lines …

Discussion

There are several steps in this slicing protocol designed to promote tissue health and favor the emergence of spontaneous naturalistic network activity: the mouse is transcardially perfused with chilled sucrose cutting solution; horizontal-entorhinal cortex (HEC) slices are cut at a thickness of 450 μm from the intermediate or ventral hippocampus; slices recover at the interface of warmed ACSF and humidified, carbogen-rich air; during recordings slices are superfused with ACSF warmed to 32 °C and delivered at a…

Disclosures

The authors have nothing to disclose.

Acknowledgements

The author would like to thank Steve Siegelbaum for support. Funding is provided by 5R01NS106983-02 as well as 1 F31 NS113466-01.

Materials

3D printer Lulzbot LulzBot TAZ 6
Acute brain slice incubation holder NIH 3D Print Exchange 3DPX-001623 Designed by ChiaMing Lee, available at https://3dprint.nih.gov/discover/3dpx-001623
Adenosine 5′-triphosphate magnesium salt Sigma Aldrich A9187-500MG
Ag-Cl ground pellets Warner 64-1309, (E205)
agar Becton, Dickinson 214530-500g
ascorbic acid Alfa Aesar 36237
beaker (250 mL) Kimax 14000-250
beaker (400 mL) Kimax 14000-400
biocytin Sigma Aldrich B4261
blender Oster BRLY07-B00-NP0
Bonn scissors, small becton, Dickinson 14184-09
borosilicate glass capillaries with filament (O.D. 1.5 mm, I.D. 0.86 mm, length 10 cm) Sutter Instruments BF150-86-10HP Fire polished capillaries are preferable.
calcium chloride solution (1 M) G-Biosciences R040
camera Olympus OLY-150
compressed carbogen gas (95% oxygen / 5% carbon dioxide) Airgas X02OX95C2003102
compressed oxygen Airgas OX 200
constant voltage isolated stimulator Digitimer Ltd. DS2A-Mk.II
coverslips (22×50 mm) VWR 16004-314
cyanoacrylate adhesive Krazy Glue KG925 Ideally use the brush-on form for precision
data acquisition software Axograph N/A Any equivalent software (e.g. pClamp) would work.
Dell Precision T1500 Tower Workstation Desktop Dell N/A Catalog number will depend on specific computer – any computer will work as long as it can run electrophysiology acquisition software.
Digidata 1440A Molecular Devices 1-2950-0367
digital timer VWR 62344-641 4-channel Traceable timer
disposable absorbant pads VWR 56616-018
dissector scissors Fine Science Tools 14082-09
double-edge razor blades Personna BP9020
dual automatic temperature controller Warner Instrument Corporation TC-344B
dual-surface or laminar-flow optimized recording chamber N/A N/A The chamber presented in this protocol is custom made. A commercial equivalent would be the RC-27L from Warner Instruments.
equipment rack Automate Scientific FR-EQ70" A rack is not strictly necessary but useful for organizing electrophysiology
Ethylene glycol-bis(2-aminoethyiether)- N,N,N',N'-teetraacetic acid (EGTA) Sigma Aldrich 324626-25GM
filter paper Whatman 1004 070
fine scale Mettler Toledo XS204DR
Flaming/Brown micropipette puller Sutter Instruments P-97
glass petri dish (100 x 15 mm) Corning 3160-101
glucose Fisher Scientific D16-1
Guanosine 5′-triphosphate sodium salt hydrate Sigma Aldrich G8877-250MG
ice buckets Sigma Aldrich BAM168072002-1EA
isoflurane vaporizer General Anesthetic Services Tec 3
lab tape Fisher Scientific 15-901-10R
lens paper Fisher Scientific 11-996
light source Olympus TH4-100
magnesium chloride solution (1 M) Quality Biological 351-033-721EA
magnetic stir bars Fisher Scientific 14-513-56 Catalog number will be dependent on the size of the stir bar.
micromanipulator Luigs & Neumann SM-5
micromanipulator (manual) Scientifica LBM-2000-00
microscope Olympus BX51WI
microspatula Fine Science Tools 10089-11
monitor Dell 2007FPb
MultiClamp 700B Microelectrode Amplifier Molecular Devices MULTICLAMP 700B The MultiClamp 700B should include headstages, pipette holders, and a model cell.
N-(2-Hydroxyethyl)piperazine-N′-(2-ethanesulfonic acid), (HEPES) Sigma Aldrich H3375-25G
needle (20 gauge, 1.5 in length) Becton, Dickinson 305176
nylon filament YLI Wonder Invisible Thread 212-15-004 size 0.004. This cat. # is from Amazon.com
nylon mesh Warner Instruments Corporation 64-0198
perstaltic pump Harvard Apparatus 70-2027
Phosphocreatine di(tris) salt Sigma Aldrich P1937-1G
pipette holders Molecular Devices 1-HL-U
platinum wire World Precision PT0203
polylactic acid (PLA) filament Ultimaker RAL 9010
potassium chloride Sigma Aldrich P3911-500G
potassium gluconate Sigma Aldrich 1550001-200MG
potassium hydroxide Sigma Aldrich 60377-1KG
razor blades VWR 55411-050
roller clamp World Precision Instruments 14041
scale Mettler Toledo PM2000
scalpel handle Fine Science Tools 10004-13
slice harp Warner SHD-26GH/2
sodium bicarbonate Fisher Chemical S233-500
sodium chloride Sigma Aldrich S9888-1KG
sodium phosphate monobasic anhydrous Fisher Chemical S369-500
sodium pyruvate Fisher Chemical BP356-100
spatula VWR 82027-520
spatula/spoon, large VWR 470149-442
sterile scalpel blades Feather 72044-10
stirrer / hot plate Corning 6795-220
stopcock valves, 1-way World Precision Instruments 14054
stopcock valves, 3-way World Precision Instruments 14036
sucrose Acros Organics AC177142500
support for swivel clamps Fisher Scientific 14-679Q
surgical scissors, sharp/blunt Fine Science Tools 14001-12
syringe (1 mL) Becton, Dickinson 309659
syringe (60 mL with Luer-Lok tip) Becton, Dickinson 309653
three-pronged clamp Fisher Scientific 05-769-8Q
tissue forceps, large Fine Science Tools 11021-15
tissue forceps, small Fine Science Tools 11023-10
transfer pipettes Fisher Scientific 13-711-7M
tubing Tygon E-3603 ID 1/16 inch, OD 3/16 inch
tubing Tygon R-3603 ID 1/8 inch, OD 1/4 inch
vacuum grease Dow Corning 14-635-5D
vibrating blade microtome Leica VT 1200S
vibration-dampening table with faraday cage Micro-G / TMC-ametek 2536-516-4-30PE
volumetric flask (1 L) Kimax KIM-28014-1000
volumetric flask (2 L) PYREX 65640-2000
warm water bath VWR 1209
 

References

  1. Buzsáki, G., Lai-Wo, S., Vanderwolf, C. H. Cellular bases of hippocampal EEG in the behaving rat. Brain Research Reviews. 6, 139-171 (1983).
  2. Buzsáki, G. Hippocampal sharp waves: Their origin and significance. Brain Research. 398, 242-253 (1986).
  3. Buzsáki, G., Horváth, Z., Urioste, R., Hetke, J., Wise, K. High-frequency network oscillation in the hippocampus. Science. 256, 1025-1027 (1992).
  4. Buzsáki, G., Anastassiou, C. A., Koch, C. The origin of extracellular fields and currents — EEG, ECoG, LFP and spikes. Nature Reviews Neuroscience. 13, 407-420 (2012).
  5. Buzsáki, G. Hippocampal sharp wave-ripple: A cognitive biomarker for episodic memory and planning. Hippocampus. 25, 1073 (2015).
  6. Maier, N., et al. Reduction of high-frequency network oscillations (ripples) and pathological network discharges in hippocampal slices from connexin 36-deficient mice. Journal of Physiology. 541, 521-528 (2002).
  7. Maier, N., Nimmrich, V., Draguhn, A. Cellular and network mechanisms underlying spontaneous sharp wave-ripple complexes in mouse hippocampal slices. Journal of Physiology. 550, 873-887 (2003).
  8. ul Haq, R., et al. Adrenergic modulation of sharp wave-ripple activity in rat hippocampal slices. Hippocampus. 22, 516-533 (2012).
  9. ul Haq, R., et al. Serotonin dependent masking of hippocampal sharp wave ripples. Neuropharmacology. 101, 188-203 (2016).
  10. Maier, P., Kaiser, M. E., Grinevich, V., Draguhn, A., Both, M. Differential effects of oxytocin on mouse hippocampal oscillations in vitro. European Journal of Neuroscience. 44, 2885-2898 (2016).
  11. Mizunuma, M., et al. Unbalanced excitability underlies offline reactivation of behaviorally activated neurons. Nature Neuroscience. 17, 503-505 (2014).
  12. Hájos, N., et al. Spike timing of distinct types of GABAergic interneuron during hippocampal gamma oscillations in vitro. Journal of Neuroscience. 24, 9127-9137 (2004).
  13. Geschwill, P., et al. Synchronicity of excitatory inputs drives hippocampal networks to distinct oscillatory patterns. Hippocampus. , (2020).
  14. Rutecki, P. A., Grossmann, R. G., Armstrong, D., Irish-Loewen, S. Electrophysiological connections between the hippocampus and entorhinal cortex in patients with complex partial seizures. Journal of Neurosurgery. 70, 667-675 (1989).
  15. Lothman, E. W., Bertram, E. H., Stringer, J. L. Functional anatomy of hippocampal seizures. Progress in Neurobiology. 37, 1-82 (1991).
  16. Carter, D. S., Deshpande, L. S., Rafiq, A., Sombati, S., Delorenzo, R. J. Characterization of spontaneous recurrent epileptiform discharges in hippocampal – cortical slices prepared from chronic epileptic animals. Seizure: European Journal of Epilepsy. 20, 218-224 (2011).
  17. Karlócai, M. R., et al. Physiological sharp wave-ripples and interictal events in vitro: What’s the difference. Brain. 137, 463-485 (2014).
  18. Leroy, F., et al. Input-timing-dependent plasticity in the hippocampal CA2 region and its potential role in social memory. Neuron. 95, 1089-1102 (2017).
  19. Sun, Q., et al. Proximodistal heterogeneity of hippocampal CA3 pyramidal neuron intrinsic properties, connectivity, and reactivation during memory recall. Neuron. 95, 656-672 (2017).
  20. Masurkar, A. V., et al. Medial and lateral entorhinal cortex differentially excite deep versus superficial CA1 pyramidal neurons. Cell Reports. 18, 1-13 (2017).
  21. Papatheodoropoulos, C., Kostopoulos, G. Spontaneous, low frequency (∼2-3 Hz) field activity generated in rat ventral hippocampal slices perfused with normal medium. Brain Research Bulletin. 57, 187-193 (2002).
  22. Papatheodoropoulos, C., Kostopoulos, G. Spontaneous GABAA-dependent synchronous periodic activity in adult rat ventral hippocampal slices. Neuroscience Letters. 319, 17-20 (2002).
  23. Kubota, D., Colgin, L. L., Casale, M., Brucher, F. A., Lynch, G. Endogenous waves in hippocampal slices. Journal of Neurophysiology. 89, 81-89 (2003).
  24. Behrens, C. J., Van Den Boom, L. P., De Hoz, L., Friedman, A., Heinemann, U. Induction of sharp wave – complexes in vitro and reorganization of hippocampal networks. Nature Neuroscience. 8, 1560-1567 (2005).
  25. Kouvaros, S., Papatheodoropoulos, C. Prominent differences in sharp waves, ripples and complex spike bursts between the dorsal and the ventral rat hippocampus. Neuroscience. 352, 131-143 (2017).
  26. Strange, B. A., Witter, M. P., Lein, E. S., Moser, E. I. Functional organization of the hippocampal longitudinal axis. Nature Reviews Neuroscience. 15, 655-669 (2014).
  27. Patel, J., Fujisawa, S., Berényi, A., Royer, S., Buzsáki, G. Traveling Theta Waves along the Entire Septotemporal Axis of the Hippocampus. Neuron. 75, 410-417 (2012).
  28. Patel, J., Schomburg, E. W., Berényi, A., Fujisawa, S., Buzsáki, G. Local generation and propagation of ripples along the septotemporal axis of the hippocampus. Journal of Neuroscience. 33, 17029-17041 (2013).
  29. Fricke, R., Cowan, W. M. An autoradiographic study of the commissural and ipsilateral hippocampo-dentate projections in the adult rat. Journal of Comparative Neurology. 181, 253-269 (1978).
  30. Ishizuka, N. O. R., Weber, J., Amaral, D. G. Organization of intrahippocampal projections originating from CA3 pyramidal cells in the rat. The Journal of Comparative Neurology. 623, 580-623 (1990).
  31. Papatheodoropoulos, C. Electrophysiological evidence for long-axis intrinsic diversification of the hippocampus. Frontiers in Bioscience – Landmark. 23, 109-145 (2018).
  32. Gilbert, M., Racine, R. J., Smith, G. K. Epileptiform burst responses in ventral vs dorsal hippocampal slices. Brain Research. 361, 389-391 (1985).
  33. Papatheodoropoulos, C., Moschovos, C., Kostopoulos, G. Greater contribution of N-methyl-D-aspartic acid receptors in ventral compared to dorsal hippocampal slices in the expression and long-term maintenance of epileptiform activity. Neuroscience. 135, 765-779 (2005).
  34. Jones, R. S. G., Heinemann, U. Synaptic and intrinsic responses of medial entorhinal cortical cells in normal and magnesium-free medium in vitro. Journal of Neurophysiology. 59, (1988).
  35. Rafiq, A., Delorenzo, R. J., Coulter, D. A. Generation and propagation of epileptiform discharges in a combined entorhinal cortex / hippocampal slice. Journal of Neurophysiology. 70, 1962-1974 (1993).
  36. Stoop, R., Pralong, E. Functional connections and epileptic spread between hippocampus, entorhinal cortex and amygdala in a modified horizontal slice preparation of the rat brain. European Journal of Neuroscience. 12, 3651-3663 (2000).
  37. Roth, F. C., Beyer, K. M., Both, M., Draguhn, A., Egorov, A. V. Downstream effects of hippocampal sharp wave ripple oscillations on medial entorhinal cortex layer V neurons in vitro. Hippocampus. 26, 1493-1508 (2016).
  38. Bertsche, A., Bruehl, C., Pietz, J., Draguhn, A. Region- and pattern-specific effects of glutamate uptake blockers on epileptiform activity in rat brain slices. Epilepsy Research. 88, 118-126 (2010).
  39. Wu, C., Shen, H., Luk, W. P., Zhang, L. A fundamental oscillatory state of isolated rodent hippocampus. Journal of Physiology. 540, 509-527 (2002).
  40. Colgin, L. L., Jia, Y., Sabatier, J. M., Lynch, G. Blockade of NMDA receptors enhances spontaneous sharp waves in rat hippocampal slices. Neuroscience Letters. 385, 46-51 (2005).
  41. Ellender, T. J., Nissen, W., Colgin, L. L., Mann, E. O., Paulsen, O. Priming of hippocampal population bursts by individual perisomatic-targeting interneurons. The Journal of Neuroscience. 30, 5979-5991 (2010).
  42. Xiong, G., Metheny, H., Johnson, B. N., Cohen, A. S. A. Comparison of different slicing planes in preservation of major hippocampal pathway fibers in the mouse. Frontiers in Neuroanatomy. 11, 1-17 (2017).
  43. Maier, N., Morris, G., Johenning, F. W., Schmitz, D. An approach for reliably investigating hippocampal sharp wave-ripples in vitro. PLoS One. 4, 6925 (2009).
  44. Schlingloff, D., Kali, S., Freund, T. F., Hajos, N., Gulyas, A. I. Mechanisms of sharp wave initiation and ripple generation. Journal of Neuroscience. 34, 11385-11398 (2014).
  45. McCloskey, D. P., Scharfman, H. E. Progressive, potassium-sensitive epileptiform activity in hippocampal area CA3 of pilocarpine-treated rats with recurrent seizures. Epilepsy Research. 97, 92-102 (2011).
  46. McMahon, L. L., Williams, J. H., Kauer, J. A. Functionally distinct groups of interneurons identified during rhythmic carbachol oscillations in hippocampus in vitro. Journal of Neuroscience. 18, 5640-5651 (1998).
  47. Pöschel, B., Heinemann, U., Draguhn, A. High frequency oscillations in the dentate gyrus of rat hippocampal slices induced by tetanic stimulation. Brain Research. 959, 320-327 (2003).
  48. Hájos, N., et al. Maintaining network activity in submerged hippocampal slices: importance of oxygen supply. European Journal of Neuroscience. 29, 319-327 (2009).
  49. Hájos, N., Mody, I. Establishing a physiological environment for visualized in vitro brain slice recordings by increasing oxygen supply and modifying aCSF content. Journal of Neuroscience Methods. 183, 107-113 (2009).
  50. Dengler, C. G., Yue, C., Takano, H., Coulter, D. A. Massively augmented hippocampal dentate granule cell activation accompanies epilepsy development. Nature Publishing Group. , 1-17 (2017).
  51. Ting, J. T., et al. Preparation of acute brain slices using an optimized N -methyl-D-glucamine protective recovery method. Journal of Visualized Experiments. 132, 1-13 (2018).
  52. Westerhof, N., Lankhaar, J. W., Westerhof, B. E. The arterial windkessel. Medical and Biological Engineering and Computing. 47, 131-141 (2009).
  53. Shi, W. X., Bunney, B. S. A small volume chamber for electrical recording from submerged brain slices and a pulse-free medium supply system using a peristalic pump. Journal of Neuroscience Methods. 35, 235-240 (1990).
check_url/61704?article_type=t

Play Video

Cite This Article
Whitebirch, A. C. Acute Mouse Brain Slicing to Investigate Spontaneous Hippocampal Network Activity. J. Vis. Exp. (162), e61704, doi:10.3791/61704 (2020).

View Video