Waiting
Login processing...

Trial ends in Request Full Access Tell Your Colleague About Jove

Neuroscience

The 6-hydroxydopamine Rat Model of Parkinson's Disease

Published: October 27, 2021 doi: 10.3791/62923

Summary

The 6-hydroxydopamine (6-OHDA) model has been used for decades to advance the understanding of Parkinson's Disease. In this protocol, we demonstrate how to perform unilateral nigrostriatal lesions in the rat by injecting 6-OHDA in the medial forebrain bundle, assess motor deficits, and predict lesions using the stepping test.

Abstract

Motor symptoms of Parkinson's disease (PD)-bradykinesia, akinesia, and tremor at rest-are consequences of the neurodegeneration of dopaminergic neurons in the substantia nigra pars compacta (SNc) and dopaminergic striatal deficit. Animal models have been widely used to simulate human pathology in the laboratory. Rodents are the most used animal models for PD due to their ease of handling and maintenance. Moreover, the anatomy and molecular, cellular, and pharmacological mechanisms of PD are similar in rodents and humans. The infusion of the neurotoxin, 6-hydroxydopamine (6-OHDA), into a medial forebrain bundle (MFB) of rats reproduces the severe destruction of dopaminergic neurons and simulates PD symptoms. This protocol demonstrates how to perform the unilateral microinjection of 6-OHDA in the MFB in a rat model of PD and shows the motor deficits induced by 6-OHDA and predicted dopaminergic lesions through the stepping test. The 6-OHDA causes significant impairment in the number of steps performed with the contralateral forelimb.

Introduction

The main neuropathological characteristics of PD are the chronic progressive neurodegeneration of dopaminergic neurons in the substantia nigra pars compacta (SNc) and the presence of Lewy bodies containing α-synuclein protein1. As SNc dopaminergic neurons project their axons into the striatum through the nigrostriatal pathway, neurodegeneration of neurons in SNc results in a dopaminergic deficit in the striatum2. The absence of dopamine in the striatum causes an imbalance in the activities of the direct and indirect motor control pathways, which is responsible for the main motor symptoms of PD: akinesia (slow movement), bradykinesia (difficulty in starting movements), muscle stiffness, and tremor at rest3,4,5.

As the molecular and physiological mechanisms involved in the onset of PD are still not fully understood, currently available principal treatments seek to alleviate the motor symptoms through pharmacotherapies, deep brain stimulation6,7, genetic therapies8, and cell transplantation9. Therefore, preclinical research is fundamental to elucidate the mechanisms involved in the onset of PD and discover new methodologies for the early diagnosis and new therapies to prevent or stop the degeneration of neurons affected by PD10.

Animal models have been widely used to simulate human pathology in the laboratory, contributing to the advancement of medicine and science11,12,13,14. However, it is essential to emphasize that the correct choice of the animal model is fundamental for the success of the study. Therefore, the animal model must be validated in three main aspects: i) face validity, in which the animal model must have the characteristics of human pathology; ii) constructive validity, in which the animal model must have a solid theoretical basis; and iii) predictive validity, in which animal models must respond to treatments in a similar way to clinical treatment.

Currently, several animals are used as animal models for PD. The main groups include mammals, such as rodents, primates, minipigs, dogs, and cats, and other groups such as drosophila and zebrafish. Rodents are the most classic animal model for PD and the most used due to their ease of handling and maintenance. In addition, the anatomy and molecular, cellular, and pharmacological mechanisms of PD are similar in rodents and humans15.

A review published by Kin and colleagues in 2019 analyzed the principal animal model methodologies used for PD in the 2000s and found that the most used animal model involved neurotoxins such as 6-hydroxydopamine (6-OHDA) and 1-methyl-4-phenyl-1,2,3,6-tetrahydropyridine (MPTP). Both neurotoxins cause mitochondrial dysregulation in dopaminergic neurons in the nigrostriatal pathway, leading to cell death16. Another widely used model involves genetic manipulation through mutation in specific genes involved in the onset of PD, causing mitochondrial dysregulation17. Neurotoxin models are commonly used to evaluate and compare therapeutics, whereas genetic models are used to study the development of preventive therapies and idiopathic PD15.

The neurotoxin MPTP was discovered to cause parkinsonism in the mid-1980s after seven patients used the substance and exhibited severe PD symptoms. In addition to the symptoms, the patients responded to treatment with L-DOPA, which made the researchers link the molecule directly to PD. After the case was published in 1986, several researchers began using MPTP in preclinical PD research18. Researchers have found that being a lipophilic molecule, MPTP can cross the blood-brain barrier (BBB) and be converted to MPP+19. This toxic substance accumulates inside neurons and causes damage to complex 1 of the mitochondrial respiratory chain, leading to the death of dopaminergic neurons20.

The 6-OHDA neurotoxin model was first used to induce the degeneration of monoamine neurons of the nigrostriatal pathway in 196821. The 6-OHDA model is commonly used to cause neurodegeneration in the nigrostriatal pathway as it is a dopamine analog and toxic for catecholamine-containing cells. After 6-OHDA enters the brain, it may be taken up by the dopamine transporter (DAT) in dopaminergic neurons, leading to degeneration of the nigrostriatal pathway22. Because 6-OHDA does not penetrate the BBB, it must be administered directly through intracerebral stereotaxic injection23. A noradrenaline reuptake inhibitor is often combined with 6-OHDA microinjection to preserve noradrenergic fibers and provide a more selective degeneration of dopaminergic neurons24.

After DAT takes up 6-OHDA, it will accumulate in the cytosol of neurons, producing reactive oxygen species (ROS) and leading to cell death15. Three different lesions models of 6-OHDA are frequently used: i) lesions to the SNc25,26; ii) lesions to the striatum27,28; iii) lesions to the MFB29,30. Lesions caused in the striatum result in a slow and retrograde degeneration of dopaminergic neurons in SNpc. In contrast, lesions caused in SNpc and MFB result in rapid and total degeneration of neurons, leading to more advanced parkinsonian symptoms31.

Unilateral or bilateral injection of 6-OHDA can cause neurodegeneration in dopaminergic neurons. 6-OHDA does not always cause severe damage to the neurons; sometimes, the injection results in partial damage, which is also used to simulate the early stages of PD32. The unilateral injection is more commonly used due to the model's ability to assess the animal's motor deficits and predict cell loss through tests such as amphetamine/apomorphine-induced rotation and the stepping test29. Bilateral injections are most used to evaluate spatial memory and recognition33.

The amphetamine/apomorphine-induced rotation test is a behavioral test commonly used to predict cell loss in the nigrostriatal pathway. It is defined as a process in which repeated administration of dopamine agonists leads to an intensification of rotational behavior in 6-OHDA-lesioned animals34. Rotational behavior consists of quantifying amphetamine-induced ipsilateral rotation or apomorphine-induced contralateral turns in unilaterally lesioned rodents. Drug-induced rotational behavior has been criticized because rotation does not correspond to PD symptoms in humans and can be affected by variables such as tolerance, sensitization, and "priming"35.

Priming is one of the most critical factors in these behavioral tests. Some cases have been reported wherein a single dose of L-DOPA led to a failure in rotational behaviors36. Additionally, another critical factor related to the combined application of the amphetamine-induced test and apomorphine-induced test for parallel use is that they measure different endpoints because of different mechanisms of action, reflecting the inactivation of different signaling mechanisms and pathways. Furthermore, the amphetamine-induced test is more accurate to measure nigrostriatal lesions above 50-60%, whereas the apomorphine-induced test is more accurate for lesions above 80%37.

The stepping test has emerged as a behavioral test that indicates deficits related to dopaminergic neuron degeneration and therapeutic effects. It enables the analysis of akinesia caused by a 6-OHDA lesion in dopaminergic neurons without a drug-induced procedure. Furthermore, the test has been well established and commonly used since 1995, when it was first described by Olsson et al.35. In 1999, Chang et al.38 also analyzed and compared the performance of rats in the stepping test with the level of degeneration caused by 6-OHDA and found that animals that performed worse in the stepping test also had a more significant degeneration of dopaminergic neurons.

The stepping test is an excellent method to predict severe dopaminergic nigrostriatal damage in 6-OHDA-lesioned rats. Evidence suggests that motor deficits appear in the contralateral forelimb of the 6-OHDA infusion during the stepping test when the degree of dopaminergic loss in SNc is >90%39. This paper describes the protocols, methodologies, and materials used to perform stereotaxic surgery for the unilateral infusion of 6-OHDA into the MFB of rats and how to predict the dopaminergic lesions caused by the toxin through the stepping test.

Subscription Required. Please recommend JoVE to your librarian.

Protocol

All procedures involving animals followed the ethical principles of the National Council for the Control of Animal Experimentation (CONCEA) and the Arouca Law (Law 11.794/2008) and were approved by the local ethics committee (CEUA-FFCLRP/USP (18.5.35.59.5).

1. Preparation of drugs

  1. Anesthesia with Ketamine/Xylazine
    NOTE: The dose of ketamine used is 70 mg/kg, and the dose of xylazine is 10 mg/kg.
    1. To prepare 1 mL of anesthetic using ketamine 100 mg/mL solution and xylazine 20 mg/mL solution, combine 0.35 mL of ketamine solution, 0.25 mL of xylazine solution, and 0.4 mL of 0.9% sterile saline solution. Administer the anesthetic solution at a final volume of 2 mL/kg.
      NOTE: Ketamine along with xylazine can produce sedation for 60-80 min. If the animal still has reflexes (e.g., hind leg pitching and/or blinking reflex), administer an additional 10% of the individual dose.
  2. Imipramine
    NOTE: The individual dose of imipramine used is 20 mg/kg.
    1. To prepare 1 mL of imipramine 20 mg/mL solution, combine 20 mg of imipramine and 1 mL of 0.9% sterile saline solution. Administer the imipramine solution at a final volume of 1 mL/kg.
  3. Meloxicam
    NOTE: The individual dose of meloxicam used is 1 mg/kg.
    1. To prepare 1 mL of meloxicam 1 mg/mL solution, combine 0.05 mL of meloxicam 2% and 0.95 mL of 0.9% sterile saline solution. Administer the meloxicam solution at a final volume of 1 mL/kg once a day for two days.
  4. Ascorbic acid 0.1%
    1. To prepare 1 mL of 0.1% ascorbic acid, combine 1 mg of ascorbic acid and 1 mL of 0.9% sterile saline solution.
  5. 6-hydroxydopamine (6-OHDA)
    NOTE: 6-OHDA is a neurotoxin used to selectively destroy dopaminergic and noradrenergic neurons in the brain. Avoid direct contact with skin and mucous membranes of the eyes, nose, and mouth. When handling 6-OHDA, wear double nitrile gloves, lab coat, disposable gown, eye protection, and surgical mask or face shield. The total infusion volume of the toxin is 4 μL/animal, and the individual amount is 10 μg of 6-OHDA/animal.
    1. To prepare 1 mL of 6-OHDA at a final concentration of 2.5 mg/mL, mix 2.5 mg of 6-OHDA and 1 mL of 0.9% saline solution containing 0.1% ascorbic acid (described above).
      NOTE: 6-OHDA is light-sensitive and degrades faster when exposed to bright light. It must be properly handled and stored in an environment protected from light. If the color of the solution is reddish, discard it.
  6. Lidocaine hydrochloride (2%)
    1. Prepare 2% lidocaine solution for local application to the animal.
      NOTE: The maximum dose that can be applied is 7 mg/kg.
  7. Poly-antibiotic suspension
    NOTE: The poly-antibiotic suspension with streptomycins and penicillins (see the Table of Materials) must be prepared at the time of application with the entire volume of diluent, whose ampoule accompanies the vial with the powder.
    1. Remove the metallic disc on the rubber stopper. Disinfect the rubber stopper with alcohol.
    2. Using a syringe with a needle of 23 G, inject the diluent into the vial. Remove the needle and shake the vial vigorously until the suspension is entirely homogenized. Inject a little air into the vial and withdraw the desired volume of suspension.
    3. Administer a deep intramuscular injection, pulling the plunger before injecting the drug to ensure that no blood vessel is reached.
      ​NOTE: The final volume of suspension to be applied is 0.5 mL/kg.

2. Preparation of materials

NOTE: Always follow instructions provided with the material safety data sheet when handling chemicals.

  1. Stereotaxic apparatus
    1. Place the stereotaxic device on a stable and clean bench with proper illumination to perform the surgery. Disinfect the apparatus with 70% ethanol.
    2. Check if the ear and incisor bars of the device are correctly aligned. Place a thermal blanket where the animal will be placed during surgery to stay warm during the procedure. Monitor the animal's temperature with an accurate rectal probe.
      NOTE: The thermal blanket should be at 37.5 °C so that the animal maintains a body temperature of 37 °C body.
  2. Microinfusion system
    1. Fill (70-80%) a Hamilton syringe (50 μL or as desired) attached to a medical-grade polyethylene microtubing and a needle with double distilled water (ddH2O) and check for leaks through the system.
    2. Pull air through the system so that a single air bubble separates the ddH2O in the syringe from the 6-OHDA solution in the microtube.
      NOTE: This procedure avoids contaminating the Hamilton syringe with 6-OHDA and allows the use of several rats on the same experimental day.
    3. Position the Hamilton syringe on the infusion pump so that it is firmly attached and the plunger of the syringe is parallel to the frame that will move to push it. Set the infusion pump to a speed of 0.5 μL/min so that the total application of 4 μL of 6-OHDA lasts for 8 min. Test the infusion system by confirming that there are no leaks and that the infusion occurs according to the previously set time and volume.
    4. Attach the needle of infusion attached to the microtube to the apparatus at the end of the stereotaxic arm and check that the needle is positioned at a 180° angle to the surface. Ensure that the needle is straight and not bent.
      NOTE: Check all the described procedures carefully because if any of the items in the infusion system do not work correctly, it may jeopardize the success of the surgery.
  3. Suture
    1. Use a sterile nylon non-absorbable suture with a 3/8 circle needle to suture the incision after surgery.
  4. Postsurgical recovery site
    1. Place a clean and sterilized housing box where animals can be monitored until fully recovered (responsive to touch and manipulation). Put a thermal blanket in the box for thermoregulation.
      ​NOTE: As thermoregulation is important, include a supplemental heat source to maintain body temperature if necessary.

3. Surgical procedure

NOTE: In this protocol, adult male Sprague-Dawley rats (200-250 g) were kept under controlled conditions of temperature (22 ± 2 °C), air exchange (15-20 exchanges/hour), and light-dark cycles (12 h/12 h), grouped in boxes with 3 or 4 animals, with free access to food and water.

  1. Weigh the animals to monitor weight changes in the days following the surgery. Calculate the dose of drugs to be administered.
  2. Administer imipramine intraperitoneally 30 min before surgery (~10-15 min before administering anesthesia), using a 27 G needle and a 1 mL syringe.
    NOTE: The imipramine will block the noradrenaline transporter (NAT) and prevent 6-OHDA uptake by noradrenergic neurons, making the lesion more selective to the dopaminergic neurons40.
  3. After 10-15 min of the administration of imipramine, administer the intraperitoneal ketamine/xylazine anesthesia using a 27 G needle and a 1 mL syringe. Wait until the animal is completely anesthetized. Verify that the animal is under deep anesthesia when the animal does not respond to hind leg pinching and does not show a blink reflex.
  4. Shave the rat's fur in the region of the head where the incision will occur.
  5. Position the rat in the stereotaxic apparatus.
    1. Position the head over the incisor bar and fix the bar 3.3 mm below the interaural line.
    2. Position the ear bars, one side at a time. Position the incisor bar and the ear bars so that the top of the skull is straight and parallel to the surface.
    3. Adjust the nose clamp and test that the head is firm and does not move to either side.
  6. Apply sterile ophthalmic ointment to the rat eyes to prevent corneas from drying out.
  7. Apply povidone-iodine to the area to be incised to disinfect the site.
  8. Apply local lidocaine for analgesia of the incision region; do not exceed 7 mg/kg.
  9. Administer the meloxicam subcutaneously using a 27 G needle and a 1 mL syringe.
    NOTE: Meloxicam is a nonsteroidal anti-inflammatory analgesic that will help the animal recover post surgery.
  10. Administer the poly-antibiotic suspension intramuscularly using a 23 G needle and a 1 mL syringe.
    NOTE: The poly-antibiotic suspension is administered as a prophylactic treatment to avoid possible bacterial infections in the postsurgery recovery.
  11. Check that the animal is in a state of deep anesthesia by checking for blink reflexes or hind limb reflexes by pinching the hind paw with tweezers.
  12. With a scalpel, make an incision of ~1.5 cm in the region where the microinjection will occur.
    NOTE: Sterile techniques are applied from this point until the wound closure.
  13. Clean the skull region with cotton swabs and cotton buds until the Bregma and Lambda can be seen. Mark the Bregma and the Lambda with a sterilized fine pen.
  14. Check that the dorsal-ventral (DV) coordinates of Bregma and Lambda are similar. If they are different, readjust the rat in the stereotaxic apparatus as the rat's head is not correctly positioned.
  15. Note down the anteroposterior (AP) and mediolateral (ML) coordinates of the Bregma.
  16. Move to the AP and ML coordinates of the right MFB according to 41: AP: -4.3 mm, ML: 1.6 mm from Bregma.
  17. Mark the region of the trepanation with a sterilized fine pen.
  18. With a sterilized drill, slowly pierce the animal's skull, taking care not to injure the dura mater.
  19. Position the microinjection needle on the dura mater and note the DV coordinates. Take a thin needle and gently rupture the dura mater. Insert the needle to the DV coordinate (8.3 mm ventral) of the MFB, where the microinjection will take place.
  20. Operate the microinjection pump to release the 6-OHDA solution into the MFB. When the microinjection is finished, check the Hamilton syringe to see if 4 μL of 6-OHDA has been injected.
    NOTE: The microinjection should last 8 min.
  21. After administration of the 6-OHDA, wait for 10 min before removing the needle to avoid backflow of the drug. Remove the microinjection needle slowly from the animal's brain.
  22. Disinfect the incision region again with povidone-iodine.
  23. Suture the incision area with ~3-4 surgical knots.
    NOTE: The knot should not be too strong or too loose.
  24. Remove the rat from the stereotaxic apparatus and place it in a clean box for recovery on the thermal blanket until the animal has fully recovered from anesthesia. Observe the animal every 15 min until it is fully awake from anesthesia.

4. Postoperative procedures

  1. Monitor the weight of the animals over the next four days after surgery. Treat them with meloxicam subcutaneously once a day for two days after surgery, adjusting the dose for each day's weight.
    NOTE: All animals should be assessed for the need for analgesics on the third day after the surgery.
  2. Check the incisions daily for at least four days to ensure they are not infected. Look for heat, swelling, pain, discharge, and redness until the incisions heal.
  3. Check the appetite and water consumption by monitoring the animal's body weight. Give wet feed to encourage the animals to eat. Observe the general body condition, attitude, and mobility daily for at least four days after surgery. Remove the sutures 7-10 days after surgery.
    ​NOTE: Animals should be euthanized if the endpoints defined in the ethical procedures are reached.

5. Stepping test

  1. Training
    NOTE: The animals should be trained for three days before the test. According to the protocol described below, training should occur twice a day, once in the morning and once in the afternoon, or with an interval of at least 2 h between sessions. Track the time using a timer.
    1. Day 1
      1. In the first session, handle the rat by holding it in gloves for ~1-2 min to allow the rat to familiarize itself with the handler/experimenter.
      2. In the second session, alternate between holding the rat for 20 s and placing it on the protocol table for 20 s. Repeat this training step for 3 min to familiarize the rat with the experimental setup for the stepping test.
    2. Day 2
      1. In the first session, place both forepaws of the rat on the protocol table by holding its hind paws and back with one hand. Tilt the rat downwards headfirst at an angle of 45° to the flat surface of the protocol table. Move horizontally on the table from end to end, allowing the rat to step on the table with both paws (cover 90 cm in 4 s). Hold the rat in gloves for 10 s, allowing it to rest; repeat this pattern for 3 min.
      2. In the second session, place one forepaw of the rat on the protocol table by holding the other forepaw back with one hand and hold the rat's back and hind paws with the other hand (see step 5.1.2.1). Move horizontally on the table from end to end in 4 s, allowing the rat to step with its free paw. Hold the rat in gloves for 10 s, allowing it to rest, and repeat with another forepaw, followed by the rest period. Repeat this pattern, alternating between the two forepaws, and rest for 3 min.
      3. Repeat the training step 3 times for 1 min each.
    3. Day 3
      1. In the first session, follow the procedure described in step 5.1.2.2 for one forepaw. Repeat with another forepaw, followed by the rest period. Repeat this pattern, alternating between the two forepaws, and rest for 3 min.
      2. In the second session, follow the procedure described in step 5.1.2.2.
  2. Test
    NOTE: The stepping test is performed before surgery, 2 and 4 weeks after stereotaxic surgery, to evaluate the akinesia of the contralateral forelimb and the possible injury caused by 6-OHDA.
    1. Hold the rat at an angle of 45° to the surface, immobilizing its hind limbs and allowing only one of the forelimbs to rest on the platform, as explained above, for day 3 of training.
    2. Drag the rat forward over a distance of 90 cm in 4 s, with the right or left paw resting on the surface.
    3. Take notes and quantify the number of forehand-adjusting steps taken with each paw in each direction.

Subscription Required. Please recommend JoVE to your librarian.

Representative Results

Dopaminergic lesion assessment
The stepping test enables the assessment of the akinesia of the anterior limb contralateral to the lesion and the selection of animals with a possible lesion of the nigrostriatal pathway induced by 6-OHDA infusion (Figure 1). The comparison of the performance of the contralateral forelimb stepping test presurgery and 2 weeks and 4 weeks after surgery revealed interaction (F2,74 = 93.63; p < 0.0001; two-way repeated-measures ANOVA) between time (pre, 2, and 4 weeks after surgery) and treatment (sham-operated and 6-OHDA-lesioned). Bonferroni's post-hoc test showed a significant decrease in the number of steps contralateral to the lesion in animals receiving 6-OHDA in the right MFB compared to the sham-operated animals at the second and fourth week after surgery (p < 0.0001) (Figure 1). The results were consistent with those of previous studies35.

It is important to note that when the dopaminergic lesion is not complete, the results of the stepping test will not reach the degree of success of the results presented in this study. A previously published study performed the stepping test and immunohistochemistry of tyrosine hydroxylase (TH) with animals with a partial dopaminergic lesion after performing surgery for microinjection of 6-OHDA following the same protocol used in this study. Their finding of a partial deficit in the stepping test (4-8 steps) is the result of a partial dopaminergic lesion of ~60% of the neurons39.

Figure 1
Figure 1: Assessment of contralateral stepping test pre- and postsurgery for unilateral infusion of 6-OHDA or vehicle into the right MFB. Data show that animals receiving 6-OHDA had a significant decrease in the number of steps with the anterior forelimb contralateral to the lesion at the second and fourth weeks after surgery (****p < 0.0001 vs. sham postsurgery; two-way repeated-measures ANOVA, Bonferroni post-hoc). Data expressed as mean ± standard error of the mean. Vehicle is 0.9% saline solution containing 0.1% ascorbic acid. Results are based on 14 animals in the sham group and 25 animals in the 6-OHDA group. Abbreviations: P = presurgery. 2 = two weeks after surgery. 4 = four weeks after surgery; 6-OHDA = 6-hydroxydopamine; MFB = medial forebrain bundle. Please click here to view a larger version of this figure.

The comparison of the performance of the ipsilateral forelimb stepping test presurgery and 2 weeks and 4 weeks after surgery did not reveal any interaction (F2,74 = 0.4492; p = 0.6399; two-way repeated-measures ANOVA) between time (pre, 2, and 4 weeks after surgery) and treatment (sham-operated and 6-OHDA-lesioned). Bonferroni's post-hoc test did not show any significant difference in the number of steps ipsilateral to the lesion in animals receiving 6-OHDA in the right MFB compared to sham animals (Figure 2).

Figure 2
Figure 2: Assessment of ipsilateral stepping test pre- and postsurgery for unilateral infusion of 6-OHDA or vehicle into the right MFB. Data show that animals receiving 6-OHDA did not significantly decrease the number of steps with the anterior forelimb ipsilateral to the lesion at the second and fourth weeks after surgery (p > 0.05 vs. sham postsurgery; two-way repeated-measures ANOVA, Bonferroni post-hoc). Data expressed as mean ± standard error of the mean. Vehicle is 0.9% saline solution containing 0.1% ascorbic acid. Results are based on 14 animals in the sham group and 25 animals in the 6-OHDA group. Abbreviations: P = presurgery. 2 = two weeks after surgery. 4 = four weeks after surgery; 6-OHDA = 6-hydroxydopamine; MFB = medial forebrain bundle. Please click here to view a larger version of this figure.

Consistent with previous studies on 6-OHDA-lesioned animals42, histological analysis (Figure 3) comparing TH of the striatum of both hemispheres allows a reliable assessment of the DA deficit in the striatum. Therefore, this behavioral protocol can be used in combination with immunohistochemical methods in studies involving experimental models of PD.

Figure 3
Figure 3: Representative images of TH labeling in the 6-OHDA experimental model of PD, including anterior striatum and substantia nigra compacta. The panoramic image demonstrates the extension of the lesion, and inset zooms depict innervation e cell bodies immunostained. (A) Image of the striatal coronal section showing a partial injury induced by 6-OHDA in the right hemisphere. (B) Image of substantia nigra and ventral tegmental area coronal section from the same animal also showing the lesion extension. (C) Image of the striatal coronal section showing a complete induced injury by 6-OHDA in the right hemisphere. (D) Image of substantia nigra and ventral tegmental area coronal section from the same animal also showing the lesion extension. Scale bar = 1.3 mm in panoramic view and 65 µm in inset zooms. Abbreviations: 6-OHDA = 6-hydroxydopamine; TH = tyrosine hydroxylase; PD = Parkinson's disease; NL = non-lesioned; L = lesioned. Please click here to view a larger version of this figure.

Subscription Required. Please recommend JoVE to your librarian.

Discussion

This paper describes a protocol for performing surgery for unilateral microinfusion of 6-OHDA in the MFB, capable of causing robust lesions in the neurons of the nigrostriatal pathway and generating akinesia in the animal. Also described is the protocol for performing the stepping test, an easily applicable and noninvasive test that can be used to prove the success of the lesions and assess forelimb akinesia. As presented in the representative results, animals receiving 6-OHDA showed a reduction in the number of adjusting steps contralateral to injury, which means that 6-OHDA-injured animals exhibit strong akinesia from 2 weeks after infusion surgery. Akinesia-the focus of several treatments for the disease-is one of the main motor symptoms of PD. The development of akinesia in an animal model is significant for preclinical studies of PD. Moreover, these results resemble those reported by Chang et al.37, who confirmed that animals presenting a lower number of steps had a higher percentage of dopaminergic neuron death by immunohistochemistry. Therefore, animals that presented a lower number of contralateral adjusting steps are more likely to have a dopaminergic injury.

Assessment of the success of the surgery and the lesions can also be confirmed by other behavioral tests such as amphetamine/apomorphine-induced rotation43, elevated body swing test (EBST), corridor test, cylinder test, tissue labeling techniques such as TH immunohistochemistry, or even quantification of dopamine in the striatum by HPLC42. Other methodologies differ in the injected dose of 6-OHDA and postsurgery time interval for behavioral assessment. A recent review43 summarizes the most recent articles using this methodology and the difference in dose, behavioral testing, and postsurgery interval between them. The model of PD induced by 6-OHDA does not mimic all the pathological processes related to the disease, such as the accumulation of Lewy bodies, but simulates the death of dopaminergic neurons of the striatal-nigral pathway. This enables the study of new therapies for the symptoms of the disease, which could lead to an improvement in the quality of life of patients affected by this disease.

Despite being the most widely used model, the 6-OHDA model has its limitations like all current PD models. The model has the disadvantage of not fully representing the molecular mechanisms involved in the pathology of the disease, such as the accumulation of alfa-synuclein proteins and the formation of Lewy bodies. The model simulates the death of dopaminergic neurons of the nigrostriatal pathway, corresponding to a late stage of the disease and leading to the onset of motor symptoms only. This makes it unsuitable for studying its natural development15,32. The 6-OHDA model described in this article is usually characterized by low mortality rates. Postsurgery recovery is crucial to prevent high mortality rates due to the union of an invasive procedure and the neurodegenerative lesion44. It is possible to reduce mortality by taking extra care during the postsurgery recovery period with nutritional supplementation, rehydration, and external temperature control45. The combination of such measures has been shown to reduce or even eliminate the mortality rate drastically30,46. A common cause of death is the insertion of the needle at the wrong coordinate in the brain. It is crucial to carefully check the coordinates during this delicate surgical procedure. This will avoid damage to other brain structures (e.g., the hypothalamus) by the needle, which can impair the animal's eating and drinking actions, leading to malnutrition and dehydration47.

Finally, it is essential to highlight that although the ketamine-xylazine anesthesia protocol is well established and used in rodent experiments48, some evidence suggests that the combination of these anesthetics may be insufficient for an extended period of surgery. Additionally, ketamine-xylazine sensitivity might vary according to different strains of mice and rats49,50. An alternative may be to induce anesthesia by isoflurane inhalation. One study demonstrated faster loss of the righting reflex with isoflurane-induced anesthesia than with ketamine-xylazine. Moreover, 60% of the rats anesthetized with ketamine-xylazine presented consecutive toe pinch reflexes during the surgical procedure, even with dose supplementation. In contrast, animals anesthetized with isoflurane presented isolated cases of tail pinch reflexes that disappeared after volume adjustment51.

Subscription Required. Please recommend JoVE to your librarian.

Disclosures

The authors have no conflicts of interest to declare.

Acknowledgments

This work was supported by São Paulo Research Foundation (FAPESP, grant 2017/00003-0). We are grateful for the Coordination for the Improvement of Higher Education Personnel (CAPES). We thank Dr. Anthony R. West, Dr. Heinz Steiner, and Dr. Kuei Y. Tseng for support and mentoring.

Materials

Name Company Catalog Number Comments
6-OHDA Sigma Aldrich H4381 https://www.sigmaaldrich.com/catalog/product/sigma/h4381?lang=pt&region=BR&cm_sp=Insite-_-caSrpResults_srpRecs_srpModel
_6-ohda-_-srpRecs3-1
70% Alcohol
Ascorbic acid Sigma Aldrich 795437 https://www.sigmaaldrich.com/catalog/product/sial/795437?lang=pt&region=BR&gclid=
Cj0KCQjw4cOEBhDMARIsAA3XD
RipyOnxOxkKAm3J1PxvIsvw09
_kfaS2jYcD9E5OyuHYr4n89kO
6yicaAot6EALw_wcB
Cotton
Drill or tap
Gauze
Hamilton syringe 50 uL Hamilton 80539 https://www.hamiltoncompany.com/laboratory-products/syringes/80539
Imipramine Alfa Aeser J63723 https://www.alfa.com/pt/catalog/J63723/
Infusion pump Insight EFF-311 https://insightltda.com.br/produto/eff-311-bomba-de-infusao-2-seringas/
Ketamine (Dopalen) Ceva https://www.ceva.com.br/Produtos/Lista-de-Produtos/DOPALEN
Machine for trichotomy
Meloxicam (Maxicam 2%  Ourofino) Ourofino https://terrazoo.com.br/produto/maxicam-injetavel-2-50ml-ouro-fino/
Metal Disposal
Paper towels
Pentabiotic Zoetis https://www.zoetis.com.br/pentabiotico-veterinario.aspx
Plastic waste garbage can
Poly-antibiotic Pentabiotic (Wealth)
Povidone-iodine
Scalpel and blades
Scissors
Scraper
Stereotaxic apparatus Insight EFF-331 https://insightltda.com.br/produto/eff-331-estereotaxico-1-torre/
Sterile saline solution
Swabs
Temperature probe
Timer
Tweezers
Xylazine (Anasedan) Ceva https://www.ceva.com.br/Produtos/Lista-de-Produtos/ANASEDAN

DOWNLOAD MATERIALS LIST

References

  1. Gibb, W. R., Lees, A. J. The relevance of the Lewy body to the pathogenesis of idiopathic Parkinson's disease. Journal of Neurology, Neurosurgery, and Psychiatry. 51 (6), 745-752 (1988).
  2. Albin, R. L., Young, A. B., Penney, J. B. The functional anatomy of basal ganglia disorders. Trends in Neurosciences. 12 (10), 366-375 (1989).
  3. Dexter, D. T., Jenner, P. Parkinson disease: from pathology to molecular disease mechanisms. Free Radical Biology & Medicine. 62, 132-144 (2013).
  4. Obeso, J. A., et al. Functional organization of the basal ganglia: therapeutic implications for Parkinson's disease. Movement Disorders. 23, Suppl 3 548-559 (2008).
  5. Tysnes, O. -B., Storstein, A. Epidemiology of Parkinson's disease. Journal of Neural Transmission. 124 (8), 901-905 (2017).
  6. Karachi, C., et al. Clinical and anatomical predictors for freezing of gait and falls after subthalamic deep brain stimulation in Parkinson's disease patients. Parkinsonism & Related Disorders. 62, 91-97 (2019).
  7. Sudhakar, V., Richardson, R. M. Gene therapy for Parkinson's disease. Progress in Neurological Surgery. 33, 253-264 (2018).
  8. Baizabal-Carvallo, J. F., et al. Combined pallidal and subthalamic nucleus deep brain stimulation in secondary dystonia-parkinsonism. Parkinsonism & Related Disorders. 19 (5), 566-568 (2013).
  9. Morizane, A. Cell therapy for Parkinson's disease with induced pluripotent stem cells. Clinical Neurology. 59 (3), 119-124 (2019).
  10. Jankovic, J., Tan, E. K. Parkinson's disease: etiopathogenesis and treatment. Journal of Neurology, Neurosurgery, and Psychiatry. 91 (8), 795-808 (2020).
  11. Cenci, M. A., Whishaw, I. Q., Schallert, T. Animal models of neurological deficits: how relevant is the rat. Nature Reviws. Neuroscience. 3 (7), 574-579 (2002).
  12. Tronci, E., Francardo, V. Animal models of l-DOPA-induced dyskinesia: the 6-OHDA-lesioned rat and mouse. Journal of Neural Transmission. 125 (8), 1137-1144 (2018).
  13. Lane, E., Dunnett, S. Animal models of Parkinson's disease and L-dopa induced dyskinesia: How close are we to the clinic. Psychopharmacology. 199 (3), 303-312 (2008).
  14. Meredith, G. E., Sonsalla, P. K., Chesselet, M. -F. Animal models of Parkinson's disease progression. Acta Neuropathologica. 115 (4), 385-398 (2008).
  15. Kin, K., Yasuhara, T., Kameda, M., Date, I. Animal models for Parkinson's disease research: trends in the 2000s. International Journal of Molecular Sciences. 20 (21), 5402 (2019).
  16. Schober, A. Classic toxin-induced animal models of Parkinson's disease: 6-OHDA and MPTP. Cell and Tissue Research. 318 (1), 215-224 (2004).
  17. Smith, G. A., Isacson, O., Dunnett, S. B. The search for genetic mouse models of prodromal Parkinson's disease. Experimental Neurology. 237 (2), 267-273 (2012).
  18. Langston, J. W., Ballard, P., Tetrud, J. W., Irwin, I. Chronic Parkinsonism in humans due to a product of meperidine-analog synthesis. Science. 219 (4587), 979-980 (1983).
  19. Langston, J. W., Irwin, I., Langston, E. B., Forno, L. S. 1-Methyl-4-phenylpyridinium ion (MPP+): identification of a metabolite of MPTP, a toxin selective to the substantia nigra. Neuroscience Letters. 48 (1), 87-92 (1984).
  20. Ramsay, R. R., Salach, J. I., Singer, T. P. Uptake of the neurotoxin 1-methyl-4-phenylpyridine (MPP+) by mitochondria and its relation to the inhibition of the mitochondrial oxidation of NAD+-linked substrates by MPP+. Biochemical and Biophysical Research Communications. 134 (2), 743-748 (1986).
  21. Ungerstedt, U. 6-Hydroxy-dopamine induced degeneration of central monoamine neurons. European Journal of Pharmacology. 5 (1), 107-110 (1968).
  22. Blandini, F., Armentero, M. -T. Animal models of Parkinson's disease. FEBS Journal. 279 (7), 1156-1166 (2012).
  23. McDowell, K., Chesselet, M. -F. Animal models of the non-motor features of Parkinson's disease. Neurobiology of Disease. 46 (3), 597-606 (2012).
  24. Luthman, J., Fredriksson, A., Sundström, E., Jonsson, G., Archer, T. Selective lesion of central dopamine or noradrenaline neuron systems in the neonatal rat: motor behavior and monoamine alterations at adult stage. Behavioural Brain Research. 33 (3), 267-277 (1989).
  25. Casarrubea, M., et al. Effects of Substantia Nigra pars compacta lesion on the behavioral sequencing in the 6-OHDA model of Parkinson's disease. Behavioural Brain Research. 362, 28-35 (2019).
  26. Wang, R., Shao, M. L-DOPA-elicited abnormal involuntary movements in the rats damaged severely in substantia nigra by 6-hydroxydopamine. Annals of Palliative Medicine. 9 (3), 947-956 (2020).
  27. Hernandez-Baltazar, D., Mendoza-Garrido, M. E., Martinez-Fong, D. Activation of GSK-3β and caspase-3 occurs in Nigral dopamine neurons during the development of apoptosis activated by a striatal injection of 6-hydroxydopamine. PLoS One. 8 (8), 70951 (2013).
  28. Bagga, V., Dunnett, S. B., Fricker, R. A. The 6-OHDA mouse model of Parkinson's disease - Terminal striatal lesions provide a superior measure of neuronal loss and replacement than median forebrain bundle lesions. Behavioural Brain Research. 288, 107-117 (2015).
  29. Iancu, R., Mohapel, P., Brundin, P., Paul, G. Behavioral characterization of a unilateral 6-OHDA-lesion model of Parkinson's disease in mice. Behavioural Brain Research. 162 (1), 1-10 (2005).
  30. Boix, J., Padel, T., Paul, G. A partial lesion model of Parkinson's disease in mice - Characterization of a 6-OHDA-induced medial forebrain bundle lesion. Behavioural Brain Research. 284, 196-206 (2015).
  31. Blesa, J., Phani, S., Jackson-Lewis, V., Przedborski, S. Classic and new animal models of Parkinson's disease. Journal of Biomedicine & Biotechnology. 2012, 845618 (2012).
  32. Breit, S., et al. Effects of 6-hydroxydopamine-induced severe or partial lesion of the nigrostriatal pathway on the neuronal activity of pallido-subthalamic network in the rat. Experimental Neurology. 205 (1), 36-47 (2007).
  33. More, S. V., Kumar, H., Cho, D. -Y., Yun, Y. -S., Choi, D. -K. Toxin-induced experimental models of learning and memory impairment. International Journal of Molecular Sciences. 17 (9), 1447 (2016).
  34. Schwarting, R. K. W., Huston, J. P. The unilateral 6-hydroxydopamine lesion model in behavioral brain research. Analysis of functional deficits, recovery and treatments. Progress in Neurobiology. 50 (2-3), 275-331 (1996).
  35. Olsson, M., Nikkhah, G., Bentlage, C., Björklund, A. Forelimb akinesia in the rat Parkinson model: differential effects of dopamine agonists and nigral transplants as assessed by a new stepping test. Journal of Neuroscience. 15 (5), 3863-3875 (1995).
  36. Lindgren, H. S., Rylander, D., Ohlin, K. E., Lundblad, M., Cenci, M. A. The 'motor complication syndrome' in rats with 6-OHDA lesions treated chronically with l-DOPA: Relation to dose and route of administration. Behavioural Brain Research. 177 (1), 150-159 (2007).
  37. Björklund, A., Dunnett, S. B. The amphetamine induced rotation test: A re-assessment of its use as a tool to monitor motor impairment and functional recovery in rodent models of Parkinson's disease. Journal of Parkinson's Disease. 9 (1), 17-29 (2019).
  38. Chang, J. W., Wachtel, S. R., Young, D., Kang, U. J. Biochemical and anatomical characterization of forepaw adjusting steps in rat models of Parkinson's disease: studies on medial forebrain bundle and striatal lesions. Neuroscience. 88 (2), 617-628 (1999).
  39. Jayasinghe, V. R., Flores-Barrera, E., West, A. R., Tseng, K. Y. Frequency-dependent corticostriatal disinhibition resulting from chronic dopamine depletion: role of local striatal cGMP and GABA-AR signaling. Cerebral Cortex. 27 (1), 625-634 (2017).
  40. Schallert, T., Fleming, S. M., Leasure, J. L., Tillerson, J. L., Bland, S. T. CNS plasticity and assessment of forelimb sensorimotor outcome in unilateral rat models of stroke, cortical ablation, parkinsonism and spinal cord injury. Neuropharmacology. 39 (5), 777-787 (2000).
  41. Paxinos, G., Watson, C. The rat brain in stereotaxic coordinates. , Academic Press. (2006).
  42. Padovan-Neto, F. E., et al. Selective regulation of 5-HT1B serotonin receptor expression in the striatum by dopamine depletion and repeated L-DOPA treatment: relationship to L-DOPA-induced dyskinesias. Molecular Neurobiology. 57 (2), 736-751 (2020).
  43. Prasad, E. M., Hung, S. -Y. Behavioral tests in neurotoxin-induced aAnimal models of Parkinson's disease. Antioxidants. 9 (10), Basel, Switzerland. 1007 (2020).
  44. Lundblad, M., Picconi, B., Lindgren, H., Cenci, M. A. A model of L-DOPA-induced dyskinesia in 6-hydroxydopamine lesioned mice: Relation to motor and cellular parameters of nigrostriatal function. Neurobiology of Disease. 16 (1), 110-123 (2004).
  45. Masini, D., et al. A guide to the generation of a 6-hydroxydopamine mouse model of Parkinson's disease for the study of non-motor symptoms. Biomedicines. 9 (6), 598 (2021).
  46. Francardo, V., et al. Impact of the lesion procedure on the profiles of motor impairment and molecular responsiveness to L-DOPA in the 6-hydroxydopamine mouse model of Parkinson's disease. Neurobiology of Disease. 42 (3), 327-340 (2011).
  47. Thiele, S. L., Warre, R., Nash, J. E. Development of a unilaterally-lesioned 6-OHDA mouse model of Parkinson's disease. Journal of Visualized Experiments: JoVE. (60), e3234 (2012).
  48. Fish, R., Danneman, P., Brown, M., Karas, A. Anesthesia and Analgesia in Laboratory Animals. , Academic Press. (2008).
  49. Buitrago, S., Martin, T. E., Tetens-Woodring, J., Belicha-Villanueva, A., Wilding, G. E. Safety and efficacy of various combinations of injectable anesthetics in BALB/c mice. Journal of the American Association for Laboratory Animal Sciences. 47 (1), 11-17 (2008).
  50. Struck, M. B., Andrutis, K. A., Ramirez, H. E., Battles, A. H. Effect of a short-term fast on ketamine-xylazine anesthesia in rats. Journal of the American Association for Laboratory Animal Sciences. 50 (3), 344-348 (2011).
  51. Jiron, J. M., et al. Comparison of isoflurane, ketamine-dexmedetomidine, and ketamine-xylazine for general anesthesia during oral procedures in rice rats (Oryzomys palustris). Journal of the American Association for Laboratory Animal Sciences. 58 (1), 40-49 (2019).

Tags

6-hydroxydopamine Rat Model Parkinson's Disease Nigrostriatal Lesions Medial Forebrain Bundle Motor Deficits Stepping Test Dopaminergic Neurons Parkinsonian Symptoms Laboratory Study Physiology Anatomy Pathological Mechanism Training Sessions Forelimbs Pretest Protocol Familiarization Experimental Setup Protocol Table Bradykinesia Akinesia Tremor At Rest
The 6-hydroxydopamine Rat Model of Parkinson's Disease
Play Video
PDF DOI DOWNLOAD MATERIALS LIST

Cite this Article

Guimarães, R. P., Ribeiro, D.More

Guimarães, R. P., Ribeiro, D. L., dos Santos, K. B., Godoy, L. D., Corrêa, M. R., Padovan-Neto, F. E. The 6-hydroxydopamine Rat Model of Parkinson's Disease. J. Vis. Exp. (176), e62923, doi:10.3791/62923 (2021).

Less
Copy Citation Download Citation Reprints and Permissions
View Video

Get cutting-edge science videos from JoVE sent straight to your inbox every month.

Waiting X
Simple Hit Counter