Abstract
Huntington’s disease (HD) is an autosomal neurodegenerative disease that is characterized by an excessive number of CAG trinucleotide repeats within the huntingtin gene (HTT). HD patients can present with a variety of symptoms including chorea, behavioural and psychiatric abnormalities and cognitive decline. Each patient has a unique combination of symptoms, and although these can be managed using a range of medications and non-drug treatments there is currently no cure for the disease. Current therapies prescribed for HD can be categorized by the symptom they treat. These categories include chorea medication, antipsychotic medication, antidepressants, mood stabilizing medication as well as non-drug therapies. Fortunately, there are also many new HD therapeutics currently undergoing clinical trials that target the disease at its origin; lowering the levels of mutant huntingtin protein (mHTT). Currently, much attention is being directed to antisense oligonucleotide (ASO) therapies, which bind to pre-RNA or mRNA and can alter protein expression via RNA degradation, blocking translation or splice modulation. Other potential therapies in clinical development include RNA interference (RNAi) therapies, RNA targeting small molecule therapies, stem cell therapies, antibody therapies, non-RNA targeting small molecule therapies and neuroinflammation targeted therapies. Potential therapies in pre-clinical development include Zinc-Finger Protein (ZFP) therapies, transcription activator-like effector nuclease (TALEN) therapies and clustered regularly interspaced short palindromic repeats (CRISPR)/CRISPR-associated system (Cas) therapies. This comprehensive review aims to discuss the efficacy of current HD treatments and explore the clinical trial progress of emerging potential HD therapeutics.
Keywords: chorea, antipsychotic medication, antidepressants, mood stabilizers, antisense oligonucleotides, RNAi therapies, Huntington’s disease
Introduction
HD is an autosomal dominant neurodegenerative disease that currently has no approved cure. HD is characterized by an increased number of CAG trinucleotide repeats in the huntingtin gene (HTT) on the short arm of chromosome 4. A higher number of CAG repeats in the huntingtin gene is directly associated with a greater probability of HD onset. Individuals with greater than 36 CAG repeats have an increased risk of HD, while patients with over 40 repeats are predestined to experience disease onset within their life span.1-5 The HTT gene which possesses the CAG repeats is transcribed into the huntingtin protein (HTT). 6 A high number of CAG repeats results in a dysfunctional mHTT protein, forming protein aggregates within the neurons and other cell types. 7 These aggregates, among other mechanisms, are hypothesized to cause mitochondrial dysfunction, neuronal stress, excitotoxicity and neuroinflammation (Figure 1).8-12 The onset of HD typically occurs at middle age, however, HD symptoms can develop at any time during the lifespan.13,14 Worldwide, the prevalence of HD is 2.71 per 100,000, however, this number increases to 5.70 per 100,000 in westernised countries such as Europe, the United States and Australia. 15 This has been postulated to be the result of a higher frequency of the mutant HD gene in western haplotypes. 16
Figure 1.
Mechanisms of Huntington’s disease pathogenesis.
A variety of symptoms have been recorded with the majority resulting in the deterioration of cognitive ability, motor function, a change in behaviour or a combination of these symptoms.17,18 At present, there are no approved treatments that directly target the mutated HTT gene or corresponding mHTT aggregates. Current approved HD drugs address the range of symptoms that arise as a consequence of the disease, but they do not address the disease at its origin. 19 Each patient’s symptoms can be unique, hence there is no standardized treatment with medication prescribed on a case-by-case basis and an extensive range of medications used to treat each of the unique symptoms.13,20 Side effects from medications used to treat HD can also affect each patient differently. 18
Most HD symptoms can be classified under motor, cognitive or behavioural categories.14,18 Motor symptoms include chorea during the early and mid-stages of disease onset, however, in later stages of disease progression movements become slower (bradykinetic) and less frequent (hypokinetic). 21 Alternatively, behavioural symptoms include depression, apathy and anxiety.22,23 In 1988, a survey of 2835 HD patients discovered that 40% of patients had depression at the time of the study, while 50% had sought clinical help for depression in the past. 24 Other less common psychiatric symptoms include obsessive-compulsive behaviour, psychosis and aggression. Memory deficits and retrogression of thought processing and organization can arise as cognitive symptoms.21,24 Other symptoms can also present, such as unintentional weight loss and sleep difficulties. 14
The field of potential future HD therapies is advancing rapidly and holds an exciting future. Currently, attention is particularly focussed on lowering mHTT through RNA and DNA targeting therapies. 25 Of these, antisense oligonucleotides are the most rapidly advancing potential therapies. 26 Currently, Tominersen has one of the largest phase 3 clinical trials, and results from this trial were expected in 2022.27,28 However, the dosing in this trial has been prematurely discontinued and the results so far are being assessed. 29 Most RNA and DNA targeting therapies are delivered invasively via intrathecal or intracranial injection. 30 Furthermore, DNA based therapies are all in pre-clinical stages and have added ethical complications due to their potential to edit germline DNA. 31 Other less invasive therapies are therefore being investigated, including stem cell therapies, antibody therapies (both mostly delivered intravenously) and small molecule therapies (delivered orally).25,32,33 All of these concepts hold promise and have their own strengths and weaknesses. Many are in clinical trials with manifest HD participants, 27 however, there is still significant development of these drugs needed until new therapies are approved and available to HD patients. Accessible new therapies will likely take many years to eventuate. 34 This comprehensive review will aim to compare the relative effectiveness of currently approved medications for HD and explore new treatment possibilities, with a special focus on those undergoing clinical trials. We intend to provide a broad overview of the field of current and potential future HD therapies as a forefront resource, rather than focussing on specific therapy types as many previous reviews have done. 35 A particular emphasis is placed on currently researched therapies, but therapies investigated in the past 5 to 10 years were included if they had been comprehensively studied or discussed in their subfield.
Current Therapeutic Options for Huntington’s Disease
Chorea Medications
The most commonly prescribed HD medications are targeted to reduce chorea (Table 1). One of the major pathological hallmarks of HD is the degeneration of the basal ganglia and particularly the striatum, which has been linked to the development of chorea. The basal ganglia contributes to the initiation of movement and suppression of unwanted movements. Refinement of movement by basal ganglia structures occur through the direct and indirect pathways.36-39 In HD, it is hypothesized that the early loss of the enkephalin containing medium spiny neurons result in reduced inhibition of the globus pallidus externus (GPe) and therefore increased inhibition of the subthalamic nucleus (STN). This in turn leads to decreased activation of the GPe and the substantia nigra pars reticulata (SNr) resulting in decreased inhibition of the thalamus. An increase in the resulting excitation of the cortex through excitatory thalmo-cortical projections can lead to excess uncontrolled movements causing chorea.37-39 Although, chorea is the hallmark symptom of HD, affective and cognitive symptoms can be observed decades before the onset of chorea. 40
Table 1.
Currently used medications for HD treatment.
| Class of Medication | Medication | Mechanism of Action |
|---|---|---|
| Chorea medication | Tetrabenazine | Depletes central monoamines by reversibly inhibiting VMAT2 |
| Deutetrabenazine | Depletes central monoamines by reversibly inhibiting VMAT2, however, has a longer half-life | |
| Antipsychotic medication | Olanzapine | Inhibits dopamine receptors, serotonin receptors, histamine receptors as well as a1-adrenergic and muscarinic receptors |
| Risperidone | Selectively inhibiting serotonin and dopamine-D2 receptors | |
| Antidepressants (SSRIs) | Citalopram | Inhibit the reuptake of 5-HT into the presynaptic nerve terminal |
| Fluoxetine | Inhibit the reuptake of 5-HT into the presynaptic nerve terminal | |
| Sertraline | Inhibit the reuptake of 5-HT into the presynaptic nerve terminal | |
| Mood stabilizers | Lamotrigine | Blockade of voltage-gated sodium ion channels, however, also suppresses excitatory neurotransmitters glutamate and aspartate |
| Carbamazepine | Blockade of voltage-gated sodium ion channels |
Tetrabenazine
Tetrabenazine (TBZ) was the first drug approved to treat Huntington’s disease-associated chorea. 41 Although the direct link between its mechanism of action and overall pharmacodynamic effect of reducing chorea is not fully understood, its pharmacological action is thought to selectively deplete central monoamines from the nerve terminal by reversibly inhibiting the human vesicular monoamine transporter 2 (VMAT2). 42 To understand the effectiveness of the drug, a randomized control trial was completed by the Huntington’s Study Group where 84 patients received TBZ (n = 54) or placebo (n = 30) over 12 weeks. 43 The dose was gradually increased over the first 7 weeks to either a maximum of 100 mg/day, until desired anti-chorea effects were achieved, or until side effects prevented continued participation of the patient. The outcome was measured via a change in chorea score of the Unified Huntington’s Disease Rating Scale (UHDRS). 44 Following the 12 weeks, it was concluded that TBZ significantly reduced chorea severity by 5.0 UHDRS units relative to only 1.5 UHDRS units in the placebo treatment. Four subjects did not complete the treatment due to adverse side effects. Following on from this, another 15-year longitudinal study examined patients taking TBZ and found that 17.2% of HD patients experienced adverse side effects including drowsiness, parkinsonism, depression, insomnia, anxiety and akathisia. However, the majority of these side effects were eliminated with a reduction in dose. 45
Deutetrabenazine
A similar medication to TBZ used to treat HD-associated chorea is deutetrabenazine (DBZ). DBZ is an isotopic isomer of TBZ, where 6 hydrogen atoms have been replaced by deuterium atoms. Therefore, DBZ has the same chemical function as TBZ, however, the deuterium atoms extend the half-life of the compound, in turn allowing for less frequent administration.20,46 A clinical trial, completed by Frank et al. (2016) randomized 90 HD patients between being administered DBZ (n = 45) or a placebo (n = 45) over a 12- week period (NCT01795859). Patient chorea scores were both measured at baseline and following the treatment intervention. HD patients receiving DBZ all experienced significant improvements in their chorea scores at the end of the 12 weeks relative to the control patients. 47
Both TBZ and DBZ prescriptions are labelled with warnings about the side effects of depression and suicidality. 48 However, individuals with HD taking DBZ had a significantly lower risk of developing neuropsychiatric adverse outcomes (agitation, depression, drowsiness, insomnia and parkinsonism) relative to individuals with HD taking tetrabenazine. 49 Since the induction of the drug in a clinical setting various studies have been completed analysing the link between suicidality and TBZ/DBZ. Experiments from both Dorsney et al. (2013), and Shultz et al. (2018), both found that TBZ use was not associated with an increased incidence of depression or suicidality.50,51 The contrast in findings could be due to the unaccounted increase in depression and suicidality from the HD disease alone. 24 It is known that depression significantly increases in HD patients with 10% to 25% attempting suicide at least once.24,52 These studies highlight the need for further research to be conducted to understand whether the increase in depression and suicidality is a result of TBZ/DBZ or if HD patients already have an underlying susceptibility to these symptoms.
Dopamine antagonists (Antipsychotics or neuroleptics)
Dopamine antagonists are used as second-line agents for chorea relief. These medications are primarily used for the treatment of psychosis, however, they also help to relieve chorea symptoms due to their interaction with the D2 receptor. Specific medications that antagonise dopamine are discussed in further detail below under antipsychotic medication53-55.
Anti-glutamatergics
There has also been some research into the role that glutamate excitotoxicity plays in HD-associated chorea. Hence, medications such as amantadine and riluzole have been trialed as anti-glutamatergics to reduce chorea symptoms. 56
Amantadine
Amantadine is a non-competitive N-methyl-D-aspartate (NMDA) receptor antagonist. However, it is hypothesized to affect the dopaminergic system and is considered the most effective agent to treat levodopa-induced dyskinesia in patients with Parkinson’s disease. 57 Verhagen Metman et al. trialed Amantadine to treat HD-associated chorea. To evaluate Amantidine’s effects, 24 HD patients entered a double blinded study where the treatment group received 400 mg/day of Amantadine. Over a 2-week treatment period, chorea scores were lower in the treatment group, with the median decrease in extremity chorea at rest being 36% (P = .04). 58
Riluzole
Riluzole is another anti-glutamatergic that has also been trialed to treat Huntington’s associated chorea. One study by Rosas et al. treated 8 HD patients with 50 mg of riluzole twice a day over 6 weeks. They found that the patients’ chorea rating score improved by 35% (P = .013) and regressed back to their original scores once discontinued (P = .026). 59 Alternatively, a similar study treated 9 HD patients with an identical dosing regime, but over 12 months. They discovered that patients had improved chorea scores at 3 months, however, this was sustained at 12 months. 60 Therefore, the long-term efficacy of riluzole to treat HD-specific chorea is still in question.
Antipsychotic Medication
Olanzapine
Olanzapine is the second most commonly prescribed drug for HD 20 (Table 1). Olanzapine is traditionally used as an antipsychotic drug that is primarily used to treat symptoms associated with schizophrenia. 61 Olanzapine is now prescribed to treat many HD patients expressing behavioural symptomology. A study on the short-term effects of olanzapine with HD patients (n = 11) found that 6 months following administration their behavioural assessment scores within the UHDRS significantly improved. 62 Chorea scores also improved in these patients, however, were not significant. This suggests that olanzapine may be a more suitable medication for HD patients with behavioural symptoms, while TBZ/DBZ would better address HD-associated chorea. Patients who present with multiple symptom groups are often prescribed multiple medications, taking into account possible negative drug-drug interactions. 21 This example again illustrates that there is no universal approach to treating HD and highlights the need for each patient to be assessed individually.
Olanzapine inhibits dopamine receptors (D1, D2 and D4), serotonin receptors (5-HT2A, 5-HT2C), histamine receptors (H1) as well as a1-adrenergic and muscarinic receptors. 54 Unlike the majority of other antipsychotic medications, olanzapine has a higher affinity for 5-HT receptors compared to D2 receptors, therefore, being classed as an atypical antipsychotic. 54
Side effects of this medication include excess weight gain and dyslipidemia. This has been observed in rat studies as well as clinical trials.63,64 To overcome this, it is important to take into account the patient’s current metabolic health before prescribing the medication. A study by Ball et al., (2001) found that by placing olanzapine patients on a weight loss program they could mitigate some of the weight gain effects. However, sustained weight management was only observed in males. 65
Risperidone
Another traditionally used antipsychotic medication that is prescribed to HD patients is risperidone. 20 Similar to olanzapine, risperidone has antagonistic properties, selectively inhibiting serotonin-S2 (5-HT) and dopamine-D2 receptors 55 and is commonly used to treat schizophrenia. 66 Risperidone is an atypical antipsychotic drug. An atypical antipsychotic is a drug that has a higher binding affinity to 5-HT receptors (e.g. olanzapine), while alternatively, a typical antipsychotic has a higher binding affinity to D2 receptors. 67 However, it is important to note that while risperidone is classified as an atypical antipsychotic drug it binds to both 5-HT2 and D2 receptors. 68
Research completed by Duff et al. (2008) administered risperidone to 17 HD patients over 14 months. They found that patients taking risperidone medication had significantly improved behavioural and psychiatric assessment scores compared to controls (n = 12). 69 Additionally, motor symptoms of the treatment group stabilized while the control group’s motor symptoms deteriorated 69 (NCT04201834). This suggests that risperidone may have dual benefits for both the treatment of motor and behavioural symptoms associated with HD. However, further evidence supporting this dual benefit of risperidone is minimal, therefore, larger randomized control trials must to be completed in the future to understand the drug’s full efficacy.
A randomized control trial was completed by Conley & Mahmoud, comparing the relative efficacy of risperidone and olanzapine at reducing adverse cognitive and mood behaviours in Schizophrenic patients (n = 377) over 8 weeks. 70 Both treatments were efficacious and well tolerated by patients. The only difference was that a greater proportion of patients taking olanzapine (27%) experienced a significant increase in body weight percentage relative to those taking risperidone (12%). 70 This suggests that weight gain is an area of concern when prescribing antipsychotic drugs to HD patients, and highlights risperidone as a potentially more suitable option.
Antidepressants
Approximately 40% of HD patients experience partial or ongoing HD-associated depression and 1 in 10 attempt suicide post-disease diagnosis.24,52,71 Antidepressant drugs are often prescribed to HD patients suffering from depression, however, these do not address other psychotic symptoms nor the Huntington’s associated chorea. The majority of antidepressants used to treat HD-associated depression are selective serotonin reuptake inhibitors (SSRIs) (Table 1). SSRIs preferentially inhibit the reuptake of 5-HT into the presynaptic nerve terminal. This increases the concentration of 5-HT in the extracellular environment, increasing the amount of 5-HT available to bind to receptors on the post-synaptic cleft,72-74 which leads to desensitization of the receptors. Desensitization of serotonin receptors may contribute to the therapeutic actions of SSRIs and also account for the development of tolerance to acute side effects of these drugs. SSRIs might result in strong but slow disinhibition of 5-HT neurotransmission. Treatment with antidepressants also causes a downregulation of 5-HT2 receptors.75-77
SSRIs also have several undesirable side effects such as sexual dysfunction and weight gain. 78 These symptoms typically arise following a month of treatment and can be prevented if diagnosed early. 78 Some studies have also found an increased risk of suicide in adolescents taking SSRIs, however, this may be less of an issue in HD patients as HD onset does not usually occur until middle age. This same risk factor has not been observed in adults using SSRI medication.79,80 Overall, the literature suggests an inadequate knowledge to properly manage how antidepressants are prescribed to HD patients. 81 Further research is required to understand which SRRI’s are most suitable for HD patients and whether these are the best medication to address the high rates of suicidality.
Citalopram
A common SSRI that is prescribed to HD patients is citalopram. A randomized control trial by Beglinger et al., (2014) administered HD patients with citalopram and measured any relative changes in executive function. Thirty-three adult HD patients were either administered with 20 mg of citalopram or a placebo pill daily over 20 weeks. Although there were no significant changes in executive function, patients showed a significant improvement in depression symptoms on the Hamilton Depression Rating Scale (HAM-D). 82
Fluoxetine
A second SSRI that is prescribed to HD patients is fluoxetine. This drug is used primarily to treat HD-associated unipolar depression. 21 One case study found that 2 non-depressed HD patients who took fluoxetine experienced reduced obsessive compulsive disorder (OCD) symptoms and mood improvements. 83 Fluoxetine acts to reduce depression in the short-term, however, there are limited clinical trials examining its effectiveness in HD patients with a need for further studies becoming evident. 84
Sertraline
Another SSRI that is often prescribed to HD patients is sertraline. This medication not only reduces depression but has also been recorded to cause a complete cessation of aggression and OCD symptoms associated with HD.85,86
Mood Stabilizing Medications
HD patients with behavioural symptoms can also present with symptoms such as OCD, aggression and bipolar disorder. To stabilize these mood disorders, patients can be prescribed a range of anticonvulsants such as sodium valproate, lamtringine and carbamazepine. 21 Anticonvulsants have been reported to be effective at minimizing the severity of these symptoms in numerous studies.87-92
Traditionally, anticonvulsants were commonly used for treating seizures. 93 Seizures are caused by synchronized neuron populations firing abnormally and excessively.94,95 Anticonvulsants reduce this synchrony via 3 main mechanisms: (1) modulation of voltage-gated ion channels; (2) upregulation of inhibitory neurotransmitters; (3) downregulation of excitatory neurotransmitters. 96 However, anticonvulsants like sodium valproate and carbamazepine are also involved in longer term mechanisms such as modulating cellular plasticity and resilience. This is achieved through regulation of inositol biosynthesis (MIP synthase), cyclic adenosine monophosphate (c-AMP) response element binding protein, brain-derived neurotrophic factor (BDNF), mitogen-activated protein kinases and arachidonic acid pathway regulation, among others. 97 Furthermore, sodium valproate inhibits histone deacetylase, creating long-term gene expression changes thought to produce long-term mood stabilizing effects. 98
The exact pharmacodynamics of anticonvulsants is specific to each drug and can include several of the aforementioned mechanisms. 96 Nonetheless, all drugs have similar effects on stabilizing mood in HD patients.21,99-101
Lamotrigine
Lamotrigine also blocks voltage-gated sodium ion channels, however, it also suppresses excitatory neurotransmitters glutamate and aspartate. 102 Although lamotrigine is used as a mood stabilizer in HD, some studies suggest its suppression of excitatory neurotransmitters may reduce neuron excitotoxicity, thus, mitigating one of the pathological symptoms in HD.21,103 A double blind, placebo-controlled study was completed on 64 early diagnosed HD patients (<5 years) over 30 months. The effectiveness of treatment was analysed using a quantified neurological examination and a set of cognitive and motor tests. Lamotrigine did not significantly slow the progression of HD over the 30 months. Therefore, although beneficial against 1 symptom, lamotrigine does not possess neuroprotective attributes.103,104
Carbamazepine
Carbamazepine’s mechanism of action is solely through the blockade of voltage-gated sodium ion channels. However, the exact pharmacodynamics of the drug has not been fully elucidated and there is still some debate within the literature around this.105,106 Although, carbamazepine is prescribed as a mood stabilizing drug for HD, no clinical research was found on its specific benefits to HD patients. 21
Anticonvulsant drugs exhibit a number of side effects. General side effects can include hypersensitivity reactions, blood dyscrasias, dizziness, gastrointestinal disturbances, depression and hyponatremia. 21 Aside from these side effects, carbamazepine also has the potential to cause serious skin conditions such as Stevens-Johnson syndrome or toxic epidermal necrolysis. 107 This could suggest why carbamazepine is prescribed less often than other anticonvulsants. 21
Drug treatment can be beneficial to HD patients, although prescribed drugs should be based on concurrent symptoms as each case is unique. Patients are required to be closely monitored when taking medications and management should be patient specific. Additionally, the side effects of each drug mentioned also need to be carefully recorded and ensured they are balanced with their benefits.
Non-Invasive Strategies and Life-Style Adaptations
The overall treatment of HD requires a holistic approach to health, involving a range of health professionals. 108 Non-invasive strategies and life-style adaptations can also play a vital role in the patient’s management of HD. Chorea is the primary symptom that can severely impair balance and gait. Physiotherapists can support patients by extending the proper function of these actions as well as assessing when walking aids or wheelchairs are required.109,110 Similarly, occupational therapists can assist HD patients through completing home assessments and installing bathroom adaptions or handrails so patients can prolong living in the comfort of their own home. 111
Patients who exhibit motor symptoms can also develop speech impairments such as hyperkinetic dysarthria. Hyperkinetic dysarthria involves involuntary muscle movements in the mouth, throat and respiratory system, resulting in voice hoarseness and abnormal prosody. 112 For these cases, patients can receive speech and language therapy to optimize their verbal communication for as long as possible. If a patient presents with an inability to speak, speech and language therapists can provide communication charts or electronic communicating devices, enabling patients to continue conversing with friends and family.113,114
Malnutrition is another major issue for HD patients, with the majority having a lower than average body weight. 115 Dieticians can help patients in these circumstances, providing them with diet plans and methods to overcome the chorea-induced eating deficits such as blending foods or providing liquid food substitutes. 115 A Percutaneous Endoscopic Gastronomy (PEG) may be performed for some patients in which swallowing is severely affected 116
To alleviate the psychological and cognitive symptoms associated with HD, sessions with a psychologist can be used alongside the medications provided. This may not only improve the patient’s mental health, but also can be used to monitor how the patient is responding to each medication they are receiving. 21
Surgical Treatments
An alternative surgical treatment is deep brain stimulation.117,118 This intervention is primarily used to improve chorea symptoms.119,120 A study from Gonzalez et al. (2014), found deep brain stimulation to be effective at reducing chorea in pharmacologically resistant chorea HD patients. It did not, however, reduce dystonia or bradykinesia. 121 As pharmacological resistant HD is rare, the use of this intervention is rare. The infrequent use of this treatment can also be attributed to the additional risks associated with the invasive procedure. 122
Overall, treatment of HD should not solely be the clinician’s responsibility and the medications they prescribe. For optimal treatment, a range of health care professionals should be involved with an extensive network of communication between them. 123
Potential Future Therapeutic Options for Huntington’s Disease
Antisense Oligonucleotide (ASO) Therapies
During the past decade, attention has moved from targeting broad spectrum and downstream processes to targeting root causes for potential HD therapies Table 2, such as lowering mHTT (Figure 1). 30 ASO therapies hold significant attention in potential HD therapeutics, with multiple ASOs currently in clinical trials. 25 ASOs are single stranded oligonucleotide analogues. 31 They bind to pre-mRNA or mRNA and can act through multiple mechanisms, including RNA degradation, blocking translation and splice modulation 124 ; these ultimately alter protein expression. ASOs can be allele-specific, indicating they specifically target mHTT, or allele non-specific, indicating they target both wild type HTT (wtHTT) and mHTT. There are 3 ASOs in clinical trials currently for HD – Tominersen (an allele non-specific ASO), WVE-120101 and WVE-120102 (allele-specific ASOs) 34 (Figure 2).
Table 2.
Overview of potential future HD treatments.
| Therapy | Sponsor | Development | Administration | Allele Specificity | Dosing Frequency | |
|---|---|---|---|---|---|---|
| RNA targeting therapies | ||||||
| ASO | ||||||
| Tominersen | Hoffmann-La Roche | Phase 3 | Intrathecal injection | Allele non-specific | Multiple doses | |
| WVE-120102/120101 | Wave life Sciences | Phase 1b/2a | Intrathecal injection | Allele-specific | Allele-specific | Multiple doses |
| WVE-120101: rs362307 (SNP1) | WVE-120102: rs362331 (SNP2) | |||||
| WVE-003 | Wave life Sciences | Pre-clinical | Undisclosed | Allele-specific: SNP3 (undisclosed) | Multiple doses | |
| (CUG)7 | BioMarin | Pre-clinical | Intracerebroventricular infusion (mice) | Allele-specific: CAG repeats | Multiple doses | |
| TTX-3360 | Triplet Therapeutics | Pre-clinical | Intracerebroventricular infusion (mice) | Not applicable | Unknown | |
| RNAi | ||||||
| AMT-130 | UniQure | Phase 1b/2a | Intrastriatal injection (delivered by AAV5) | Allele non-specific | One dose | |
| AAV.shHD2.1 a | Spark | Pre-clinical | Intracranial injection (delivered by AAV1) | Allele non-specific | One dose | |
| VY-HTT01 | Voyager | Pre-clinical | Intracranial injection (delivered by AAV1) | Allele non-specific | One dose | |
| Small molecules | ||||||
| Branaplam | Novartis Pharmaceutical | Pre-clinical | Oral | Allele non-specific | Multiple doses (weekly) | |
| PTC518 | PTC Therapeutics | Phase 1 | Oral | Allele non-specific | Multiple doses | |
| Unnamed small molecule | Nuredis | Pre-clinical | Gene deletion in animal model, intracerebroventricular bolus injection | Allele-specific: elongation cofactors required for expanded CAG repeat transcription | Single dose | |
| DNA targeting therapies | ||||||
| Zinc protein fingers | ||||||
| TAK-686 | Takeda and Sangamo | Pre-clinical | Intrastriatal injection | Allele-specific: expanded CAG repeats | Single dose | |
| ZF-KOX1 | European Research Council (undertaken by Imperial College London/Fingers4Cure) | Pre-clinical | Intraventricular injection | Allele-specific: expanded CAG repeat | Single dose | |
| CRISPR/Cas9 | ||||||
| Unnamed CRISPR/Cas9 | Harvard University | Pre-clinical | NA (in cell lines) – theoretically intracranial injection | Allele-specific: SNPs related to mHTT | Single dose | |
| Unnamed CRISPR/Cas9 | NIH and National Natural Science Foundation of China (undertaken by Emory University) | Pre-clinical | Intrastriatal injection | Allele non-specific | Single dose | |
| Stem cell therapies | ||||||
| Cellavita HD | Azidus Brasil | Phase 2/3 | Intravenous infusion | Not applicable | Multiple doses | |
| Foetal stem cell transplant | Health and Care Research Wales (undertaken by Cardiff University) | Phase 1 | Intrastriatal injection | Not applicable | Single dose | |
| Autologous stem/stromal cells | Regeneris Medical | NA (in clinical trial) | Intravenous injection | Not applicable | Unknown | |
| Antibody therapies | ||||||
| ANX005 | Annexon, Inc | Phase 3 | Intravenous infusion | Not applicable | Multiple doses | |
| VX15/2503 | Vaccinex, Inc | Phase 2 | Intravenous infusion | Not applicable | Multiple doses | |
| W20 | National Natural Science Foundation of China, National Science and Technology Major Projects of New Drugs | Pre-clinical | Intracerebroventricular injection (mice) | Allele-specific: does not impact wtHTT | Multiple doses | |
| INT41 | Vybion Inc | Pre-clinical | Intrastriatal injection (mice) | Allele-specific: binds to mHTT fragments | Single dose | |
| C6-17 | AFFiRiS | Pre-clinical | Unknown | Allele-specific – binds to HTT protein near the aa586 caspase-6 cleavage region | Unknown | |
| Other small molecule therapies | ||||||
| Pridopidine | Prilenia Therapeutics | Phase 3 | Oral | Not applicable | Multiple doses | |
| Laquinimod | Teva Pharmaceutical Industries Ltd | Phase 2 (in HD) | Oral | Not applicable | Multiple doses | |
| Fenofibrate | University of California, Irvine | Phase 2 | Oral | Not applicable | Multiple doses | |
| Neflamapimod | EIP Pharma Inc | Phase 2 | Oral | Not applicable | Multiple doses | |
| Nilotinib | Georgetown University | Phase 1 | Oral | Not applicable | Multiple doses | |
| SRX246 | Azevan Pharmaceuticals’ | Phase 2 | Oral | Not applicable | Multiple doses | |
| Varenicline | University of Auckland and University of Otago | Open label study | Oral | Not applicable | Multiple doses | |
| SAGE-718 | Sage Therapeutics | Phase 1 | Oral | Not applicable | Multiple doses | |
| PBT2 | Prana Biotechnology Limited | Phase 2 | Oral | Not applicable | Multiple doses | |
aAAV.shHD2.1 is also known as AAV.shHD2.
Figure 2.
Tominersen and WVE-120101/WVE-120102 therapy development timeline.
Tominersen, WVE-120101 and WVE-120102 function by complementary base pairing to target pre-mRNA, inducing RNase H1-mediated degradation. They reduce mHTT mRNA levels, therefore, lowering mHTT protein synthesis. Administration occurs via intrathecal injection into the cerebrospinal fluid (CSF) via lumbar puncture, allowing distribution to the central nervous system.125,126
Tominersen
Hoffman-La Roche’s RO7234292 or Tominersen (previously called IONIS-HTTRx/RG6042; IONIS Pharmaceuticals) is an ASO that binds to wild type HTT (wtHTT) and mHTT mRNA, inducing degradation. 126 It is currently the most developed HD ASO therapeutic. 27
Tominersen distributed well in mice and non-human primate brains during pre-clinical experiments. It lowered mHTT mRNA and mHTT protein in a dose-dependent, sustained fashion. Tominersen reversed the HD phenotype, improved survival and reduced brain atrophy in mice. 127 These experiments informed the following phase 1b/2a clinical trial design.
In the initial randomized, double blinded phase 1b/2a clinical trial (NCT02519036), participants (n = 46) were divided into 5 groups of ascending doses and a placebo group. Tominersen was given every 28 days in 4 doses, and participants were followed up after 4 months. 128 No serious adverse events occurred, and minor adverse event incidence was similar for Tominersen and placebo groups. CSF mHTT was reduced by 40% on average at 90 mg and 120 mg doses and Tominersen showed dose dependency.129,130
Multiple trials have commenced since NCT02519036 began. A randomized phase 2 open label extension (OLE) trial (NCT03342053) further assessed adverse events, recruiting people (n = 46) that had participated in the NCT02519036 trial.131,132 Participants were given monthly or bimonthly Tominersen 120 mg doses for 15 months. 129 The trial concluded in 2019, with results yet to be published.
GENERATION HD1 (NCT03761849) is a randomized, double blinded phase 3 trial to measure HD phenotype changes using the Composite Unified Huntington’s Disease Rating Scale (cUHDRS) and total functional capacity (UHDRS-TFC), among other measurements. Researchers will divide participants (n = 909) with manifest HD into 3 groups – placebo, and Tominersen doses (120 mg) every 8 or 16 weeks. Dosing was planned to occur for 25 months, and follow up would have been 29 months.28,133 GENERATION HD1 was expected to conclude in 2022, 28 however, was suspended upon independent data review in early 2021 due to Tominersen’s benefit to risk ratio over time. No new safety signals arose during this review. 29
A randomized phase 3 OLE clinical trial (NCT03842969/GEN-EXTEND) including previous Tominersen trial participants (N = 1100) will measure long-term tolerability, behaviour changes and adverse event incidence.131,134 Participants will receive Tominersen every 8 or 16 weeks, depending on the dose timing of their previous trial. It was expected to conclude in June 2024, 134 but has been placed on hold due to another trial’s benefit to risk ratio. 29
GEN-PEAK (NCT04000594) is a non-randomized, open label phase 1 trial investigating Tominersen’s pharmacokinetics. It will conclude in 2021. Participants (N = 20 in total) in 3 groups of ascending doses will receive 2 doses 28 days apart. CSF Tominersen and mHTT concentrations, and plasma Tominersen concentrations will be measured. 135
WVE-120101 and WVE-120102
Wave Life Sciences’ WVE-120101 and WVE-120102 are allele-specific, stereopure ASOs. They target single nucleotide polymorphisms (SNPs) specific to HD genotypes – WVE-120101 targets rs362307 (SNP1), and WVE-120102 targets rs362331 (SNP2). 136 They are allele-specific, which is thought to potentially avoid long-term negative effects since they do not lower wtHTT. 137 As they target SNPs, WVE-120101 and WVE-120102 cannot treat all HD patients – but combined use could target 80% of European HD patients. 125 Unpublished pre-clinical experiments showed they reduced mHTT mRNA without significantly altering wtHTT mRNA and protein in patient derived cell lines. 138 This led to randomized, double blinded phase 1b/2a clinical trials (PRECISION-HD1 and PRECISION-HD2; NCT03225833, NCT03225846), which recruited participants with early manifest HD and rs362307 (SNP1) or rs362331 (SNP2). Participants (n = 60 per trial) were divided into 6 groups – 2 mg, 4 mg, 8 mg, 16 mg or 32 mg WVE-120101/WVE-120102 doses or placebo (.9% NaCl).126,136 The parallel studies intended to measure adverse events, tolerability, pharmacokinetics and pharmacodynamics. 136
Preliminary PRECISION-HD2 results showed WVE-120102 reduced CSF mHTT protein by 12.4% compared to placebo across all groups (n = 44). This was statistically significant in higher doses, and indicated dose dependency. Total HTT did not change between groups. There were no serious adverse events, and less minor and moderate adverse events for WVE-120102 (72%) than placebo (83%). 126 However, in early 2021 Wave Life Sciences announced PRECISION-HD2 trial data showed WVE-120102 caused no change in mHTT compared to placebo. PRECISION-HD1 data showed similar results, therefore, both trials were stopped. 139
OLE trials began for patients who participated in the PRECISION trials (NCT04617847 for WVE-120101, NCT04617860 for WVE-120102). They received multiple 8 mg or 16 mg doses of WVE-120101/WVE-120102.126,138 OLE results for PRECISION-HD2 indicated slight but inconsistent mHTT lowering, without the ability to consistently lower mHTT with increased doses. Dosing in OLEs has stopped for WVE-120101 and WVE-120102 given their inability to significantly lower mHTT and therefore be of clinical benefit. 139
Wave Life Science’s WVE-003 is a pre-clinical ASO that targets an unspecified SNP (SNP3) in mHTT mRNA. 140 WVE-003 reduced mHTT mRNA in vitro and achieved mHTT mRNA knockdown in vivo. 141 Application for a clinical trial was submitted in late 2020. 140
Other ASOs
BioMarin’s (CUG)7 is a pre-clinical allele-specific ASO which targets expanded CAG repeats in mHTT mRNA. 31 After promising results in HD patient derived fibroblasts, 142 researchers found CUG (7) reduced mHTT protein concentration in R6/2 and Q175 HD model mice by 15-60%, improved motor symptoms and increased cortical volume. 143 R6/2 mice are transgenic HD models created by expressing exon one of the human mHTT gene, containing 150 CAG repeats. 144 Q175 HD knock-in mice express human mHTT within the mouse HD gene, containing 179 CAG repeats. 145
A number of further pre-clinical HD ASO therapeutics are currently in development. Some groups are investigating ASOs that target SNPs associated with HD alleles, therefore, aiming to selectively suppress mHTT expression.146,147 Results in HD mice model neurons, 146 and in vivo using HD model mice 147 were positive. Total HTT reducing and mHTT specific ASOs have improved cognitive and behavioural deficits in HD model mice with safety and tolerability, and reduced HTT in the cortex and limbic structures in non-human primate brains (where cognitive and behavioural symptoms largely originate). 148 Another ASO induces exon 12 skipping in HTT pre-mRNA, preventing cleavage sites being expressed and therefore 586 amino acid N-terminal huntingtin fragments being created. Rodent studies showed successful exon skipping, a shorter HTT protein and reduced 586 amino acid N-terminal huntingtin fragments. 149
Some ASOs are non–HTT-targeting, such as Triplet Therapeutics’ TTX-3360. It targets MSH3 within the DNA Damage Response pathway, which is often involved in repeat expansion. TTX-3360 is safe and tolerable in HD model mice, and 50% MSH3 knockdown reduced or inhibited mHTT repeat expansion. An Investigational New Drug/Clinical Trial Application submission is planned for late 2021. 150
RNA Interference (RNAi) Therapies
RNAi therapies use transgenes that express RNA molecules such as micro-RNA (miRNA), short hairpin RNA (shRNA) and short interfering RNA (siRNA).26,31 Cellular machinery processes these molecules, and they label their target mRNA for degradation. Degradation occurs via enzymes in the RNA-induced silencing complex (RISC), 31 which stops mRNA translation and therefore protein synthesis. 26 RNAi therapies are delivered by a viral vector via intracranial injection. 31 There are advantages and disadvantages associated with RNAi therapies. For example, shRNA therapies can result in difficulty controlling shRNA expression, but shRNA therapies also tend to produce less off target effects than other RNAi therapies and last longer.151-153 These factors, and other factors associated with miRNA and siRNA, must be considered when developing RNAi therapies.
AMT-130
AMT-130 (UniQure Biopharma B.V., named AAV5-miHTT in pre-clinical studies) is a RNAi based gene therapy. It is the only gene therapy currently in clinical trials for HD. 27 AMT-130 contains a gene encoding for a miRNA, which is delivered via adeno-associated virus vector serotype 5 (AAV5). AMT-130 activates RISC, which stops mHTT (and wtHTT) protein synthesis. AMT-130 is administered via intrastriatal injection. 26
In an HD rat model, Hu128/21 HD mice and HD patient derived neurons, AAV5-miHTT reduced mHTT mRNA and protein levels.154-156 Hu128/21 HD mice result from crossing an HD mouse model, YAC128, with a wtHTT model, BAC21, giving them 2 full-length human HTT transcripts that are heterozygous for the HD mutation. 157 Potent mHTT suppression was sustained for 7 months post-injection in Hu128/21 HD mice. 158 In Q175 HD mice, AAV5-miHTT lowered human HTT protein and striatal (39%) and cortical mHTT aggregates 12 months post-injection. R6/2 HD mice showed motor symptom improvement and prolonged survival. 159 mHTT aggregates were suppressed in an HD rat model – reducing neuronal dysfunction. 155 AAV5-miHTT lowered human mHTT mRNA (72.8%) and protein (85.3%) in the minipig brain, and indicated dose dependency. 160 AAV5-miHTT did not create off target effects caused by gene dysregulation or incorrect processing in HD patient derived neurons, 156 but at high doses caused astrogliosis in Hu128/21 mice. 158 AMT-130 showed safety, tolerability and widespread distribution in the brain of non-human primates. 161
Phase 1b/2a clinical trials will test AMT-130’s safety and proof of concept in 26 manifest HD patients (NCT04120493). Some will be given a low or high dose via MRI guided bilateral intrastriatal injection, and others will be given placebo (imitation surgery). 27 Researchers are measuring adverse effects, CSF biomarkers, CSF AMT-130, mHTT concentration and clinical scale changes (like UHDRS) during the core study period and follow up. 162 Data from the initial 6 months of dosing in 2 patients, and 90 day data from the next 2 patients indicated AMT-130 was safe. 163 This prompted complete enrolment of the first dosing cohort (n = 10). The study is expected to be completed in 2026, with an OLE beginning in late 2021. 164
Other RNAi therapies
Voyager Therapeutics’ VY-HTT01 is an allele non-specific RNAi based gene therapy. An miRNA is expressed and delivered via adeno-associated virus vector serotype 1 (AAV1), targeting HTT mRNA degradation. It is administered via intracranial injection. 31 In YAC128 transgenic HD mice, which express full-length human HTT with 128 CAG repeats, VY-HTT01 reduced striatal HTT protein by 40%, partially corrected motor and behavioural symptoms, reduced HTT aggregates and showed effective striatal distribution.165,166 Unpublished results showed VY-HTT01 lowered striatal (67%) and thalamic (73%) HTT mRNA in non-human primates. 167 Clinical trial applications were submitted in late 2020, but are on hold to resolve chemistry, manufacturing and control issues. 168
Spark Therapeutics are also developing an RNAi based drug, using an AAV1 delivered shRNA (AAV.shHD2.1)31. Rodent studies showed motor symptom improvement, reduced human mHTT expression 169 and neuroprotection on administration with AAV.shHD2.1. 170 In non-human primates, lowering HTT expression via RNAi did not lead to neuronal degeneration or an immune response for 6 weeks post-injection into the putamen, strengthening RNAi as a potential HD therapy. 171 Clinical trials have not yet been initiated. 34
RNA Targeting Small Molecule Therapies
Many small molecule therapies alter pre-mRNA to include or exclude target exons, leading to an altered protein (also called splice modulation). They are administered orally, which is advantageous compared to other invasive administration methods. However, this also increases the risk of adverse effects due to their systemic distribution and lack of specificity. 25 All HD RNA targeting small molecule therapies are either pre-clinical or in phase 1 clinical trials currently, with no clinical trial data yet published. 34
PTC Therapeutics’ PTC518 is a RNA splice modulator that induces a psuedoexon inclusion containing a premature stop codon. This enhances mHTT mRNA degradation – resulting in lowered mHTT protein. Unpublished pre-clinical data showed PTC518 is systemically distributed (including widespread central nervous system distribution) and indicated dose dependency. In HD patients’ isolated cells PTC518 lowered mHTT mRNA and protein, and in transgenic BACHD HD mice, which express the full-length human mHTT with 97 mixed CAA-CAG repeats within exon 1, PTC518 systemically lowered HTT protein.172-174 A phase 1 trial has commenced with healthy participants, using ascending single and multiple doses. It is expected to indicate correct dosing, proof of concept and safety. Proof of concept was shown in healthy volunteers with dose-dependent HTT mRNA splicing. It was also well tolerated and showed predictable pharmacology.173,175
Novartis Pharmaceuticals’ Branaplam (also called LMI070) is an oral, weekly administered small molecule RNA splice modulator. It was initially designed to treat spinal muscular atrophy (SMA), 176 but researchers found it lowered mHTT protein in SMA and HD animal models. It also lowered HTT mRNA in HD models. A phase 2b clinical trial with manifest HD participants is planned for in 2021. 177
Nuredis are developing an RNA splice modulator that specifically targets transcription elongation cofactors such as Spt4.34,178 These are required for expanded CAG repeat transcription. Therefore, Spt4 inhibition would result in mHTT protein reduction. Pre-clinical data showed Spt4 inhibition using small molecules reduced expanded polyQ proteins in neurons, while restoring neuronal function and regional enzymatic activity. SUPT4H (a mammalian Spt4 ortholog) also reduced mHTT protein, toxicity and mHTT aggregates in neurons. 178 In R6/2 HD mice, SUPT4H reduction prolonged survival, reduced motor symptoms and reduced mHTT aggregation. Further studies conducted in zQ175 HD mice showed SUPT4H reduced mHTT mRNA and protein. 179
There are other groups that are searching for and developing RNA targeting small molecules to treat HD – including molecules modulating protein disulfide isomerase in the endoplasmic reticulum to reduce mHTT toxicity, 180 molecules altering degradation or autophagy mechanisms 181 and molecules targeting expanded CAG repeat RNA to reduce mHTT protein synthesis.182-184 CAG lengths can be somatically unstable in different tissue types. The mechanism underpinning this is unknown. Understanding and targeting this somatic instability may be a pathway to future therapies. 185
Zinc-Finger Protein (ZFP) Therapies
DNA targeting therapies affect the mHTT gene itself by introducing protein coding sequences to the brain parenchyma that bind to DNA regions. They require a viral vector for delivery and are administered intracranially. 31 ZFPs are synthetically made sequences that mimic naturally occurring motifs. They contain DNA binding components that often include multiple zinc finger peptides, which can bind to 3 to 5 target DNA nucleotides. The major ZFPs currently being developed for HD bind to expanded CAG repeats, preventing the mHTT gene’s transcription. This lowers mHTT protein levels without altering the gene itself. 25 However, zinc finger nucleases (ZFNs) can alter genetic material, allowing for the mHTT gene to be corrected or disrupted. 34 ZFPs have successfully reduced mHTT protein by 60% in the R6/2 HD mice model brain, reduced aggregation and improved motor symptoms. 186
There are 2 major ZFPs in development for HD – TAK-686 and ZF-KOX1.186-188 These therapies are both in pre-clinical stages currently. 25 TAK-686, which targets expanded CAG repeats, has shown success in reducing mHTT in knock-in HdhQ50 mice containing an HD allele with 50 CAG repeats. It also reduced mHTT by 62% and improved behavioural symptoms in R6/2 HD mice.188,189 A modified version of ZF-KOX1 (for mouse compatibility) delayed HD symptoms and inhibited mHTT expression by 77% in HD model mice. 187
Transcription Activator-like Effector Nuclease (TALEN) Therapies
TALENS have DNA binding domains that contain repeating peptides which bind to DNA nucleotides. 34 They cause double strand breaks via artificial nucleases, allowing for segment deletion or correction. 190 All TALENs in development for HD therapy are in pre-clinical or drug discovery stages. Their impacts on phenotype, toxicity and an effective delivery system have not been established yet. TALENs also have not been investigated in HD animal models. 191 However, TALENs have successfully removed expanded CAG repeats in yeast cells by inducing double strand breaks, resulting in gene conversion. They did not increase mutation rate or cause genomic rearrangement. 192
Lowered mHTT gene expression and protein aggregation was also achieved in HD patient derived fibroblasts using a TALEN and SNP specific transcription activator-like effector (TALE-SNP). The TALE-SNP prevented transcription of the mHTT gene by targeting SNPs associated with it, leaving wtHTT gene expression unaltered. TALENs were used for gene correction by creating double strand breaks which were repaired – deleting the expanded CAG repeats. Results showed that some of the TALE-SNPs reduced mHTT expression by up to 20% without altering normal HTT expression, indicating proof of concept.25,191
Clustered Regularly Interspaced Short Palindromic Repeats (CRISPR)/CRISPR-Associated System (Cas) Therapies
Cas9 (CRISPR associated protein 9) is a naturally occurring immune enzyme found in some bacteria, which has been utilized as a genome editing tool for the past decade. It functions via a single-guide RNA (sgRNA) binding to its target DNA, and Cas9 nucleases inducing double strand breaks. 34 Double strand breaks are repaired by error-prone non-homologous end joining (NHEJ), causing frameshifts which can impair gene expression.190,193 All CRISPR approaches are delivered with viral vectors and intracranial injection. 34 In HD therapeutics, CRISPR/Cas9 therapies may remove expanded CAG repeats to correct HD alleles, inactivate HD-associated alleles or target the HTT gene itself, non-specifically reducing HTT. Multiple groups are investigating CRISPR pre-clinically for HD.193-198
CRISPR/Cas9’s efficiency has been assessed in cell lines. A paired Cas9 Nickase strategy successfully excised expanded CAG repeats in exon 1 of the HTT gene in 3 distinct HD patient derived fibroblast lines, inactivating HTT. Nickases cause single strand breaks, reducing off target effects due to increased specificity. Each cell line treated with Cas9 Nickases had different CAG repeat lengths and showed approximately 70% reduction in HTT protein. 193 A strategy using CRISPR-Cas9 and piggyBac transposon was also successful in HD human induced pluripotent stem cells, correcting the mHTT allele and amending HD mediated phenotypic defects. 194
A PAM-altering, mHTT associated SNP based CRISPR/Cas9 approach is also in development. This therapy induces a deletion mutation that inactivates the mHTT allele, leaving an intact wtHTT allele. In HD patient derived cell lines, excision of CAG repeats associated with mHTT alleles was successful. 195
Some groups have investigated CRISPR in HD animal models. Deletion of the mHTT polyQ domain using an allele non-specific CRISPR/Cas9 based strategy successfully repressed human mHTT expression in HD140Q-knock-in mice’s striatum, which express full-length human mHTT with exon 1 containing 140 CAG repeats. CRISPR was administered via striatal injection. This approach reduced mHTT expression, mHTT aggregates, astrocyte reactivity and improved motor symptoms. 196 Another group injected SpCas9 or sgRNAs delivered by recombinant AAVs (abbreviated to rAAV.SpCas9 and rAAV.sgHD1/i3, respectively) into the right hemisphere of BACHD mice. Human HTT exon 1 was deleted, reducing human mHTT mRNA by 40%. The left hemisphere was used as a control, and did not demonstrate lowered mHTT expression. 198 Another group investigated in vivo striatal injection of CRISPR/Cas9 therapies (AAV1-SaCas9-HTT to target human HTT, or AAV1-SaCas9-mRosa26 as a control) in R6/2 HD mice. They found AAV1-SaCas9-HTT had reduced mHTT protein inclusions by 40%, reduced total mHTT protein by 50%, and it improved lifespan and motor symptoms. 197
Stem Cell Therapies
Stem cell therapies are another candidate for future HD therapeutics (Figure 3). Their main advantages are their potential ability to replace neurons lost due to HD pathology, improve regeneration and provide pro-survival factors. Furthermore, theoretically some long-term benefits would remain due to stem cell incorporation into the patient’s brain microenvironment. However, there are also concerns these therapies may induce immune reactions or rejection. 199
Figure 3.
Cellavita HD, foetal striatal stem CRT and autologous stromal/adipose-derived stem cell therapy development timeline.
Cellavita HD
Cellavita HD is the most developed stem cell therapy for HD currently 27 (Figure 3). It is a mesenchymal stem cell therapy, and is administered via intravenous injection. 200
Azidus Brasil began a non-randomized, open label phase 1 clinical trial in 2017 to test safety and effectiveness of their stem cell therapy for HD, with completion expected in 2023 (NCT02728115/SAVE-DH).27,200 Participants (n = 6 total) with manifest HD will be given 3 low or high monthly Cellavita HD doses. They will be followed up for 5 years to track adverse events and efficacy through the UHDRS scale and inflammatory markers, among other measurements. 200
A randomized, triple blind phase 2 clinical trial began in 2018 to test Cellavita HD’s dose-response relationship (NCT03252535/ADORE-DH).27,201 Thirty-five participants with manifest HD will be given 9 total low or high doses of Cellavita HD or placebo. Researchers are primarily measuring UHDRS to assess effective dosing. This trial will conclude in 2022. 201
An OLE phase 2/3 trial has also been registered, but is not yet recruiting (NCT04219241/ADORE-EXT). It will assess longer term safety and efficacy in 35 manifest HD participants who took part in the ADORE-DH trial.27,202 They will receive multiple doses over 24 months. UHDRS, TFC, neurological image improvement and clinical worsening will be measured, among other factors. It will conclude in 2022. 202
Foetal striatal stem cell replacement therapy
Foetal striatal stem cell replacement therapy was first investigated in HD patients over 20 years ago. Multiple early studies showed foetal striatal cell replacement therapy (CRT) was safe but relatively ineffective at altering patient symptoms. A phase 1, randomized controlled trial using unilateral intrastriatal grafts in 4 HD patients indicated no adverse events but no HD symptom improvement (ISRCTN36485475). 203 Similarly, a later trial using bilateral foetal striatal CRT grafts showed the UHDRS score was similar across those with (n = 5) and without (n = 12) striatal grafts. The grafts did not cause significant adverse events. 204
Some trials have shown HD symptom improvements. In 5 HD patients, asynchronous bilateral intrastriatal transplants of foetal striatal neuroblasts created increased striatal metabolic activity, cognitive and motor improvements in 3 of the 5 patients (compared to 22 control HD patients). However, the other 2 patients’ condition declined similarly to the control group. 205 Another group also found mild short-term symptom improvement in 6 patients, but this only persisted in 3 and did not alter overall HD progression. 206
A randomized, open label, phase 1 clinical trial using a relatively higher CRT dose to previous studies was initiated in 2017. CRT will be administered to 5 HD patients via intrastriatal injection and patients will receive immunosuppressant medication (ISRCTN52651778/TRIDENT).27,207 TRIDENT aims to assess safety at 4 weeks, and the effectiveness of an increased CRT dose. The trial will conclude in 2023. 207
However, foetal striatal CRT also raises concerns. Immunosuppressants used to reduce graft rejection have been problematic due to side effects and lowered compliance. 208 Immune responses can occur due to foetal striatal stem cell grafts – over a decade after the ISRCTN36485475 trial, 1 of the 4 patients showed strong microglial activation, immune cell infiltration and mHTT protein aggregates at the graft site. 209 However, while some groups believe immunosuppressants are required, 210 others have shown patients’ immune responses after CRT grafts are unpredictable – therefore excessive immunosuppressant therapy is undesirable. 211 Some groups have found that striatal foetal neural grafts can only provide clinical improvement for a few years, after which patients decline. Therefore, CRT may not be viable as a permanent cure. 212 Ethical concerns also persist due to the use of foetal cells, and graft survival is uncertain due to findings showing reduced vascularization, and lowered astrocytes and pericyte numbers. 213
Autologous stem cell therapies
Autologous stromal cells are also being investigated by Regeneris Medical to treat a range of neurological disorders in a current clinical trial (n = 300), including HD (NCT03297177). 27 Autologous stem cell transplantation requires reprogramming mutated cells or isolating cells with the wild type allele. 214 Pre-clinical data for autologous adipose derived stem cells transplanted bilaterally into the striatum were positive in R6/2 HD model mice. This approach reduced striatal apoptosis, prolonged lifespan, reduced mHTT aggregates and modulated striatal neuronal loss. 215
A non-randomized, placebo-controlled study will assess the safety and change in neurological function of those given autologous stem/stromal cells compared to normal saline for up to 5 years. 27 Autologous cells will be collected from patients’ subdermal fat, isolated, concentrated and neutralized before being given intravenously. The study is expected to be completed in 2023. 216
Mesenchymal stem cell therapies
Mesenchymal stem cells (MSCs) have also been trialled in HD animal models. MSCs originate in the umbilical cord (UC), amniotic fluid, bone marrow and adipose tissue. They are self-renewing, can release neuroprotective factors like BDNF and can differentiate into multiple cell types, potentially including neural cells.217,218 No teratomas or overgrowths have been found in MSC animal model studies. 217 Striatal injection of mice UC-derived MSCs in R6/2 HD mice improved spatial memory and reduced neuropathological deficits, but did not change motor impairments. 219 Striatal transplantation of UC-derived rat MSCs in 3-nitropropionic acid (3-NP) lesioned HD rats improved motor coordination, reduced striatal atrophy, prevented oxidative stress in PC12 cells and enhanced cell viability. 220 3-NP HD rats were systemically, subcutaneously administered with 3-NP, a mitochondrial toxin, to produce striatal lesions. 221 Another 3-NP HD rat study found injection of bone marrow derived rat MSCs resulted in no significant striatal inflammation, no lateral ventricle enlargement, reduced neurological and behavioural deficits and the presence of striatal neurotrophic factors. 222
Induced pluripotent stem cell-derived neural stem cells
Induced pluripotent stem cell-derived neural stem cells (iPS-NSCs) are self-renewing, reprogrammed somatic adult cells that can differentiate into many cell types. 223 Striatal injection of mice iPS-NSCs in an HD rat model was trialled and monitored using serial 18F-FDG PET/CT. Researchers found that the rats’ striatal volume was partially recovered, memory, learning and glucose metabolism improved and iPS-NSCs differentiated into striatal neurons and glia. 224 Quinolinic acid HD rats undergo striatal injections of quinolinic acid to induce an HD phenotype. 225 Another study in a quinolinic acid-induced HD rat model found striatal transplantation with human iPS-NSCs reduced behavioural deficits, reduced inflammation, promoted neurogenesis and replaced some neural cells. 226 Striatal injection with control mice iPS-NSCs into YAC128 HD mice resulted in motor improvement, increased striatal BDNF and iPS-NSC differentiation into neural cells. 227
Antibody Therapies
Antibody-based therapies are not only being evaluated for a range of tauopathies and synucleinopathies, but are also emerging as a potential therapy for monogenic disorders of the central nervous system, including HD (Figure 4).
Figure 4.
ANX005 and VX15/2503 antibody therapy development timeline.
ANX005
Annexon, Inc’s ANX005 is a monoclonal antibody that targets C1q, preventing classical complement cascade initiation. 228 When neuronal stress is already implicated during HD pathogenesis, C1q′s role in synaptic pruning can be disrupted – affecting synapse loss, neurodegeneration and neuroinflammation. Through C1q inhibition, ANX005 is thought to reduce synapse loss and neurodegeneration.229,230 ANX005 was initially intended to treat Guillain-Barré Syndrome (GBS) and Alzheimer’s Disease (AD). 228
Pre-clinical studies showed high affinity binding of ANX005 to C1q in multiple species, and successful CSF and serum C1q reduction in monkeys and rats. Complement cascade inhibition in vivo and in vitro in GBS and AD mice models was successful. Safe dosing frequency, dose dependency and tolerability was also established in monkeys and rats. 228 A murine version of ANX005 reduced neurofilament light chain and synapse loss in HD animal models. 229
A randomized, placebo-controlled, double blinded phase 1 clinical trial began in late 2016 in 27 healthy volunteers via an ascending dose schedule (NCT03010046). ANX005 was given intravenously, either on its own or with intravenous immunoglobulin (IVIg). The study aimed to assess safety, pharmacokinetics and pharmacodynamics. It was terminated in 2018 since there were no adverse events at the maximum dose. 231 Subsequently, investigators began testing ANX005 for HD, GBS, Amyotrophic Lateral Sclerosis (ALS) and Warm Autoimmune Haemolytic Anaemia (wAIHA) patients.232-235
A phase 1b clinical trial in 31 GBS patients showed promising top-line results using an ascending dose schedule. Its aims were to test safety, tolerability and pharmacokinetics. The ANX005 dose was increased after safety was established in up to 100 mg/kg in 19 participants. No significant adverse events occurred, and ANX005 successfully inhibited C1q. ANX005 was associated with early muscle strength improvement, reduced neurofilament light chain and small GBS disability scale improvements. 236 ANX005 is also in a phase 1b, open label clinical trial to assess safety in 12 GBS patients (NCT04035135), 232 and a randomized, double blinded phase 2/3 trial in 180 GBS patients to assess efficacy and safety (NCT04701164). 237
There are also 2 phase 2, open label studies in progress for ALS and wAIHA (NCT04569435, NCT04691570, respectively). ANX005 will be given to 24 participants with ALS for 12 weeks (NCT04569435), and given to 12 patients with wAIHA for 2 doses (NCT04691570). Both studies will test ANX005’s safety, tolerability, pharmacodynamics and pharmacokinetics. The ALS trial will end in late 2021, and the wAIHA trial in early 2022.234,235
An open label phase 2a clinical trial will evaluate ANX005’s safety, tolerability, pharmacokinetics and pharmacodynamics in HD patients or patients at risk for HD (n = 24) (NCT04514367). Seven ANX005 infusions will be given for up to 10 weeks, then patients will be followed for 3 months. The study will be completed in late 2021. 233
VX15/2503
Vaccinex’s VX15/2503 (also called Pepinemab) is an IgG4 monoclonal antibody that inhibits protein semaphorin 4D (SEMA4D). SEMA4D is a protein associated with neuroinflammation by promoting glial cell activation and activating signalling cascades causing oligodendrocyte and neural precursor cell apoptosis, among other processes. 238
Pre-clinical studies showed that VX15/2503 did not induce toxicity or adverse events in rats or non-human primates. It lowered T-cell SEMA4D and induced long-term saturation with 5 weekly doses. Dose dependency was established. 238 Anti-SEMA4D is not immunosuppressive and improves blood brain barrier integrity. 239 Another study found that anti-SEMA4D improved cortical and striatal atrophy, and cognitive symptoms in YAC128 HD model mice. It did not influence motor symptoms. 240
Vaccinex, Inc and the Huntington Study Group conducted a double blinded, phase 2 clinical trial testing VX15/2503’s safety, tolerability and pharmacokinetics in late prodromal and early manifest HD patients (NCT02481674/SIGNAL). Patients were divided into 2 cohorts, A (12 months treatment) and B (36 months treatment), and given intravenous monthly VX15/2503 doses or a placebo. Investigators assessed MRI brain volume, motor and cognitive function and CSF biomarkers, among other measures. The study ended in mid-2020.241,242
Preliminary SIGNAL results from Cohort A (n = 36) showed VX15/2503 was tolerable and safe. Reduced metabolic activity and brain volume reduction appeared to be prevented. Cohort A were given VX15/2503 for another 5 months (open label), with a 3 month follow up. 243 However, results from cohort B1 with early manifest HD (n = 179) found that VX15/2503 did not prevent brain volume and metabolic activity reduction, or significantly alter motor, cognitive or behavioural function. 244 Trends towards cognitive benefits were found, and post-hoc analysis using Clinical Global Impression of Change in an advanced HD subpopulation confirmed this. However, these findings require further investigation. 245
VX15/2503 is also in clinical trials for AD, head and neck cancer and colorectal cancer,246-249 and has been tested in Advanced Solid Tumours, Multiple Sclerosis (MS), melanoma, non-small cell lung cancer.250-252 Advanced Solid Tumour (NCT01313065) (N = 42) and MS (NCT01764737) (N = 10) trials found VX15/2503 was safe and tolerable.250,251
Anti-mHTT antibody therapies
Anti-mHTT antibodies have been tested pre-clinically for over 20 years. Their intracellular targets, such as the PolyQ, PolyP and N-Terminal exon 1 domains, are affected using intrabodies. 32 Intrabodies are simpler forms of antibodies delivered via viral vectors that are able to block and inactivate proteins, enhance degradation and stop protein misfolding intracellularly. 253
PolyQ targeting intrabodies have had negative results, but PolyP/proline-rich region and N-Terminal exon 1 domain targeting intrabodies have had mixed results. rAAV6-INT41 (developed by Vybion Inc) is a PolyP/proline-rich region targeting intrabody that can lower gene dysregulation caused by mHTT.32,254 It was delivered in R6/2 HD mice by a viral vector (rAAV6) via bilateral intrastriatal injection. INT41 can reduce small mHTT aggregates by 31% and larger aggregates by 16%, while improving cognitive deficits in R6/2 HD mice. 254
Intrabodies targeting mHTT aggregates have had positive results, but raise concerns since conformational antibodies can also increase aggregation. 32 W20 is an oligomer specific intrabody that targets mHTT aggregates, among other aggregates, such as -synuclein and amyloid beta.32,255 BACHD mice given W20 via intracerebroventricular injection showed mHTT aggregate reduction in the striatum (by 49.5%) and cortex (by 26%), and lowered apoptotic cells in the striatum (by 73.3%) and cortex (by 69.6%). W20 lowered neuroinflammation, proinflammatory cytokine levels and reactive oxygen species, while improving motor and memory deficits. W20 does not affect WT HTT or surveilling microglia or astrocytes. 255
Other intrabodies have also been tested pre-clinically, such as HAPP1 and VL12.3. These target the P-rich epitope of the mHTT protein, and the N-terminal 17 amino acids, respectively. They were found to lower HTT exon 1 toxicity and aggregation. 256 C4 also targets the N-Terminal of HTT fragments. 257 When fused with mouse orthenine decarboxylase including a deleted C-terminal PEST motif (mODC-PEST), HD exon 1 fragments are reduced, along with lowered HD exon 1 fragment toxicity, aggregation and intrabody degradation. 258
Arguably antibody therapies’ main advantage is their ability to target extracellular mHTT, since most other therapies cannot do so. This is achieved via active (self-produced antibodies) or passive (pre-made antibodies) immunization. 259 Active immunization’s functional benefits are not yet known, and raise safety concerns due to increased immune gene dysregulation. Passive immunization has decreased mHTT in YAC128 mice. 32 Furthermore, AFFiRiS′ mAB C6-17 effectively bound and demonstrated specificity to human mHTT in cell-based assays and immunohistochemistry. C6-17 targets a site within the HTT protein near the aa586 caspase-6 cleavage region, and has blocked extracellular mHTT uptake. Distribution of C6-17 was quick and widespread in HD model mice, showed reduced mHTT in the central nervous system and improved motor symptoms.260,261
Antibody therapies may be most beneficial in combination with RNA based therapies to target both intracellular and extracellular mHTT. 259
Other Small Molecule Therapies
Pridopidine
Pridopidine (also named TV-7820; formerly named ACR16 or Huntexil) is an oral small molecule developed by Prilenia Therapeutics (previously developed by Teva Pharmaceutical Industries Ltd) for HD motor symptoms 262 (Figure 5). It is a dopamine stabilizer that binds to dopamine type 2 receptors, therefore, affecting striatal dopamine pathways involved in motor HD pathophysiology.262,263 It also activates the sigma-1 receptor, increasing BDNF production and supporting neuroprotection.262,264 Pre-clinical studies in R6/2 HD mice showed pridopidine prevented apoptosis, increased pro-survival molecules and reduced mHTT aggregate size – indicating neuroprotection. It also improved motor function. 265 Pridopidine prevented medium spiny neuron loss in YAC128 HD mice. 266
Figure 5.
Non-RNA targeting small molecule (Pridopidine, Laquinimod, VX-745 and Nilotinib) therapy development timeline.
Pridopidine was first clinically investigated as an HD therapy in 58 HD patients in a randomized, double blinded trial (ACR16C007). It showed tolerability and safety. While cognitive scores did not differ between placebo and pridopidine, voluntary motor performance improved using pridopidine. 267
Further trials assessed efficacy and safety of pridopidine. In a 26-week, double blinded phase 3 clinical trial (NCT00665223/MermaiHD), 437 HD patients were randomized to 45 mg or 90 mg pridopidine, or placebo. Modest motor improvements were found using UHDRS-total motor score (TMS), but this was a tertiary endpoint and other improvements (like the mMS) were not statistically significant. 268
A 12-week double blinded phase 2/3 clinical trial (NCT00724048/HART) randomized 227 HD patients to 3 dose groups or placebo. No statistically significant modified motor score (mMS) improvements were found, but trends pointed towards voluntary motor function improvement at 90 mg. 269 Both studies found acceptable tolerability and safety. Similarly, a 26-week, double blinded, phase 2 clinical trial found tolerability (NCT02006472/PRIDE-HD). 408 HD patients were randomized to 4 differing pridopidine dose groups or placebo. Investigators found no UHDRS change between placebo and pridopidine groups. A few serious adverse events occurred that likely related to HD symptoms, and 1 death during the trial was possibly affected by pridopidine administration.263,270
The PRIDE-HD trial was extended from 26 to 52 weeks in 262 patients. At 52 weeks, those taking 45 mg of Pridopidine had a lowered TFC compared to placebo, indicating slowed HD progression. Patients administered 45 mg did not experience TFC decline for the entire year they were taking pridopidine. They also found Pridopidine also has a dose-specific effect. 271
These findings led to an OLE for the MermaiHD trial and for the HART trial (NCT01306929/OPEN-HART). Both studies found long-term tolerability and safety.272,273 In OPEN-HART, 225 HD patients were administered 45 mg Pridopidine twice daily over 36 months. The trial found HD progressed similarly between placebo and Pridopidine groups for the first 36 months (as measured by TFC and TMS). 273 However, additional data showed that while TFC declined over the full 60 months, the rate of decline slowed for those taking pridopidine compared to placebo. TFC and TMS also remained stable between 48 and 60 months. However, the sample size had dropped to 33 patients by 60 months which could create confounding due to those remaining in the trial potentially having a less aggressive disease form. 274 An OLE study was also initiated for PRIDE-HD participants (NCT02494778/OPEN PRIDE-HD) 275
A double blinded, phase 3 clinical trial (NCT04556656/PROOF-HD) will investigate efficacy (via UHDRS-TFC score) and safety in 480 HD patients. Participants will be randomized to 45 mg Pridopidine twice daily or a placebo. The study will end in 2023 and an OLE is planned for PROOF-HD participants. 276
Laquinimod
Laquinimod, developed by Teva Pharmaceutical Industries, is an orally administered immunomodulator therapy 277 (Figure 5). It likely reduces central nervous system leukocyte infiltration and is neuroprotective by stimulating BDNF production, but its precise mechanism of action is unclear. 278 Pre-clinical studies showed laquinimod decreases microglial activation, reactive astrogliosis, astrocytic NF- B activation, proinflammatory cytokine production, proinflammatory changes and oligodendroglial apoptosis, among other neuroinflammatory processes. 279 Laquinimod has been tested as an MS therapeutic for over 15 years, 278 with phase 3 clinical trials showing positive efficacy results and confirming safety and tolerability (NCT00509145/ALLEGRO). 280 YAC128 mice treated with laquinimod displayed improvements in motor function and motor learning and further, displayed a reduction in depressive-like behaviours suggesting it a possible therapy for HD patients. 281
Laquinimod was tested in a randomized, double blinded, 52-week phase 2 clinical trial assessing efficacy (via UHDRS-TMS and other clinical measurements), safety and caudate volume atrophy (NCT02215616/LEGATO-HD). Participants (N = 352 total) with HD received 0.5 mg, 1 mg or 1.5 mg laquinimod or placebo.277,282 The 1.5 mg dose was terminated in 2016 due to safety concerns prompted by a study on laquinimod for MS. 277 LEGATO-HD did not show a statistically significant improvement in UHDRS-TMS, but caudate volume significantly improved. Laquinimod was safe and well tolerated. 283 However, further analysis showed that astrocytosis and gliosis was reduced in laquinimod groups. 284 Some Q-motor measurements were also statistically significantly improved in 0.5 mg laquinimod, however, multiplicity corrections were not performed. 285
VX-745
EIP Pharma’s VX-745 (also called Neflamapimod) modulates neuroinflammation and microglial dysfunction, and decreases microglial proinflammatory cytokine upregulation via p38α MAPK inhibition.286,287 VX-745 has been investigated for over 20 years, 288 but has only recently been trialled for HD 27 (Figure 5). In AD model mice, VX-745 reduced hippocampal interleukin-1 beta (IL-1 ) and improved cognitive performance – but these benefits’ peaks occurred at different doses, indicating independent processes were occurring. 287
EIP Pharma Inc and Voisin Consulting, Inc will fund a randomized, placebo-controlled, double blinded, 10-week phase 2 clinical trial in patients with early HD using orally administered VX-745 (n = 16) (NCT03980938). 27 It will primarily assess hippocampal dysfunction changes after 20 weeks. The study began in 2019 and is currently recruiting. 289
VX-745 treatment has also been assessed in AD and Lewy Body Dementia (LBD). A trial in LBD showed cognition and mobility benefits with VX-745 administration, 290 but a study on mild AD did not indicate statistically significant clinical benefit. 291 All studies to date have shown VX-745 is safe and tolerable.290,291
Nilotinib
Nilotinib (also called Tasigna® and AMN107) is being investigated as an oral HD therapeutic27,292 (Figure 5). It is a BCR-ABL tyrosine kinase inhibitor that traditionally treats chronic myeloid leukaemia (CML). 292 Nilotinib reduced oxidative stress in Parkinson’s Disease (PD) model mice via ABL inhibition and improved motor performance. 293 Tyrosine kinase inhibitors also modulated misfolded protein pathophysiology by autophagic clearance in PD model mice, positively affecting neuroinflammation. 294 Clinical trials with PD and LBD participants (n = 12) used lower Nilotinib doses than those that treat CML. Investigators found Nilotinib was safe and tolerable, and appeared to inhibit central nervous system ABL. However, the study was small and not placebo-controlled, necessitating further trials. 292 Georgetown University are recruiting for an open label phase 1 clinical trial to assess safety, tolerability and CSF biomarkers associated with motor and behavioural HD symptoms (n = 20) (NCT03764215/Tasigna-HD).27,295 Ten participants with HD will be given 150 mg for 3 months, and if safe and tolerable another ten participants will be given 300 mg for a further 3 months. There is no control group. 295
PF-02545920
Multiple therapies have been discontinued despite promising pre-clinical results. Pfizer’s PF-02545920 is an oral phosphodiesterase 10a (PDE10a) inhibitor.277,296 It aims to treat HD motor symptoms. PDE10a inhibition improves basal ganglia pathology associated with HD, such as improving spiny projection neuron pathology and responsiveness in HD mice. 297 A randomized, double blinded, 26-week phase 2 trial assessing safety and efficacy (via UHDRS-TMS) of PF-02545920 compared 5 mg and 20 mg PF-02545920 with placebo in 272 HD patients with chorea (NCT02197130/Amaryllis).277,296 Results showed no change in UHDRS-TMS or secondary endpoints between placebo and PF-02545920, despite PF-02545920 being safe and tolerable. An OLE study for Amaryllis participants was terminated after these results, 277 and PF-02545920 was discontinued from Pfizer’s drug development pipeline. 298
Other small molecule therapies are also being investigated. Fenofibrate is a peroxisome proliferator-activated receptor (PPAR) agonist, 27 capable of modifying γ coactivator-1α (PGC-1α). 299 PGC-1α is a transcriptional co-regulator involved in mitochondrial malfunction in HD, leading to neuronal degeneration. 300 Fenofibrate is in a 6 month, randomized, placebo-controlled, double blinded, phase 2 clinical trial assessing safety and efficacy (NCT03515213). 27 Azevan Pharmaceuticals’ SRX246 is a vasopressin 1a receptor antagonist that has completed Phase 2 clinical trials (NCT02507284), aimed at treating HD patients with irritability and aggression. It showed safety, tolerability and validity as a treatment candidate in the near future. 301 Varenicline is a nicotinic acetylcholine receptor agonist targeting cognitive improvements in early HD patients who smoke. HD patients treated with varenicline showed significant executive function and emotional recognition improvements, without clinically significant adverse events. 302 Sage Therapeutics’ SAGE-718 is an NMDA receptor positive allosteric modulator (PAM) that targets HD cognitive symptoms. Phase 1 results showed tolerability, no clinically significant adverse events and improved cognitive performance from baseline, leading to planned phase 2 trial initiation for late 2021.303,304 PBT2 likely reduces metal induced mHTT aggregation through metal protein attenuation. It has completed phase 2 trials (NCT01590888), showing tolerability and safety but no clear benefit for cognitive scores – warranting further studies. 305
Further Neuroinflammation Targeted Therapies for HD
Minocycline
The presence of chronic inflammation in HD has presented a potential therapeutic target with a number of treatments targeting inflammation being proposed. The second-generation tetracycline drug, minocycline, was 1 such proposed therapy that showed anti-inflammatory and anti-apoptotic properties. Mice treated with a low dose of minocycline brought about better behavioural performance than HD mice without treatment (5 mg/kg/day). 306 Stabilization of general neuropsychological and motor symptoms along with significant improvements in psychiatric symptoms were seen in patients administered with 100 mg/day of minocycline in a 2 day clinical trial. 307 However, a low sample size of 11 patients suggests that these results must be corroborated with similar studies on a larger number of patients. A further clinical trial of minocycline treatment on 30 patients over a 6 month period reported a trend toward symptom improvement on the URHDS. 308 The lack of statistical significance in these results were attributed to the short overall treatment period with the study suggesting the need for a period longer than 6 months for effective minocycline treatment. A larger study of 87 HD patients over 18 months assessed the potential benefits of this drug treatment in a phase 3 clinical trial (clinicaltrials.gov ID: NCT00277355) and found that patient improvement did not reach a futility threshold. 309 Based on these findings, further trials of minocycline administration at 200 mg/day were not warranted. 309
Cannabidiol
Cannabinoids have also been postulated to be potential therapeutic targets due to their anti-inflammatory properties.310-312 A clinical trial of cannabidiol treatment on 15 HD patients showed no improvement in motor function over a 6-week period. 313 The low sample size of this study again poses an issue when concluding the effectiveness of cannabidiol as a therapeutic agent. However, a double-blind, randomized, placebo-controlled, cross-over pilot clinical trial with Sativex (®), a botanical extract with an equimolecular combination of delta-9-tetrahydrocannabinol and cannabidiol, conducted over 12 weeks on 24 HD patients, also reported no significant symptomatic effects (clinicaltrials.gov ID: NCT01502046). 314
The failure of these clinical trials may suggest that global immunosuppression (for example through corticosteroids) or the effects of anti-inflammatory treatments may be too broad and may also highlight a need to consider the heterogeneity of microglial and astrocytic phenotypes.315-317 A coherent understanding of the diverse characteristics of these cells and their functions in a healthy state, different disease states, infection, trauma, development and ageing is critical for better therapeutic outcomes.
Conclusions
Positive steps are being made in the development of current HD therapeutics. Guidance for clinicians on HD treatment is improving – guidelines were released relatively recently to elevate international standards of HD care. 116 Current therapies focus on symptom management via multidisciplinary teams, using pharmacological and non-pharmacological treatment. 318 However, HD management has not significantly advanced for the past 20 years. 52 Furthermore, diverse and changeable symptom presentations creates complexity when treating HD. 18 There is no cure for HD, or therapeutic to alter disease onset or progression. 116 HD management requires a tailored approach not only due to symptom diversity, but also adverse effects, common contraindications, drug-drug interactions and certain HD medications worsening other symptoms.18,52,319 Clinicians must navigate these issues in addition to effectively treating HD symptoms.
Perhaps the most prominent shortcoming of today’s HD therapeutics and previous experimental therapeutics is their inability to impact specific HD targets, instead influencing non-specific and downstream processes. 30 This limits their efficiency as disease mechanisms will still propagate and cause harm where they are not inhibited. Pursuing HD-specific targets at the root cause of disease is therefore the new and likely future research direction. 30 The recent surge in interest to develop therapies that target mHTT RNA and DNA shows this trend. 31 Additional therapeutic targets are being investigated, including dopaminergic and glutamatergic neurotransmission, synaptic activity, BDNF signalling, mitochondrial function and biogenics, and aggregate inhibition.320-330 However, several of these targets were primarily researched over 10 years ago, and do not appear to be areas of significant interest in the field currently. Peripheral targets are also being investigated due to peripheral HD symptoms and HTT expression.331-333 For example, the gut microbiome is a key target due to its bidirectional communication with the brain and alteration in HD.334-338
RNA therapies, particularly antisense oligonucleotides, have favourable results and are currently of the most promising future HD therapeutics.26,339 However, RNA targeting therapies come with their own challenges. It is unknown whether non-selective HTT lowering is safe, if off target effects will occur in larger populations, whether mHTT is the singular cause of HD pathology and if permanent HTT suppression is required. Furthermore, if allele-specific treatments are required, 1 SNP will not treat all HD patients.137,340 Many RNA based therapies also require invasive administration. 31 Whether this is tolerable and effective in the long term for larger populations will likely be determined by future research. Initial clinical findings from RNAi therapies and RNA targeting small molecules, and further clinical results from antisense oligonucleotides will clarify RNA based therapies’ efficacy for HD patients. Particularly, results of larger trials such as PRECISION-HD1 and PRECISION-HD2, and further information on GENERATION-HD1 will likely inform whether antisense oligonucleotides remain the frontrunners of future HD therapeutics.
DNA therapies may also enter clinical trials within the next decade, but much work still needs to be done to achieve this goal. Ensuring DNA therapies, particularly CRISPR therapies, are safe for human use will be of primary importance. 137 Combination therapies may also be explored in future research. Current research suggests combining multiple RNA targeting therapies, 341 or RNA targeting therapies with antibodies or stem cells32,214 may be beneficial.
Work on future HD therapeutics will continue for numerous years to come and resulting treatments for HD patients are likely some time away. 34 However, many of the latest developments are encouraging and a new era of HD therapeutics may be approaching.
Footnotes
Declaration of Conflicting Interests: The author(s) declared no potential conflicts of interest with respect to the research, authorship, and/or publication of this article.
Funding: The author(s) disclosed receipt of the following financial support for the research, authorship, and/or publication of this article: This work was supported by Oakley Mental Health Research Foundation (AK, MF, TP, HJW, RLMF; 3725204), Alzheimers New Zealand Charitable Trust (AK; 370836), Alzheimers New Zealand (AK; 3718869), Freemasons New Zealand (AK; 3719321), Aotearoa Foundation, Centre for Brain Research and University of Auckland (AK; 3705579), Neurological Foundation of New Zealand (AK, TP; 848010), Brain Research New Zealand (AK, HJW, RLF), Health Research Council of New Zealand (RLF and HJW; 3627373).
ORCID iD
Andrea Kwakowsky https://orcid.org/0000-0002-3801-4956
References
- 1.Reiner A, Dragatsis I, Dietrich P. Genetics and neuropathology of Huntington's disease. Int Rev Neurobiol. 2011;98:325-372. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 2.Gusella JF, Wexler NS, Conneally PM, Naylor SL, Anderson MA, Tanzi RE, et al. A polymorphic DNA marker genetically linked to Huntington's disease. Nature. 1983;306(5940):234-238. [DOI] [PubMed] [Google Scholar]
- 3.Nance MA. Genetics of Huntington disease. Handb Clin Neurol. 2017;144:3-14. [DOI] [PubMed] [Google Scholar]
- 4.Duyao M, Ambrose C, Myers R, Novelletto A, Persichetti F, Frontali M, et al. Trinucleotide repeat length instability and age of onset in Huntington's disease. Nat Genet. 1993;4(4):387-392. [DOI] [PubMed] [Google Scholar]
- 5.The Huntington's Disease Collaborative Research Group . A novel gene containing a trinucleotide repeat that is expanded and unstable on Huntington's disease chromosomes. The Huntington's disease collaborative research group. Cell. 1993;72(6):971-983. [DOI] [PubMed] [Google Scholar]
- 6.Hoffner G, Djian P. Protein aggregation in Huntington's disease. Biochimie. 2002;84(4):273-278. [DOI] [PubMed] [Google Scholar]
- 7.Scherzinger E, Sittler A, Schweiger K, Heiser V, Lurz R, Hasenbank R, et al. Self-assembly of polyglutamine-containing huntingtin fragments into amyloid-like fibrils: Implications for Huntington’s disease pathology. Proc Natl Acad Sci Unit States Am. 1999;96(8):4604-4609. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 8.Kumar A, Ratan RR. Oxidative stress and Huntington’s disease: The good, the bad, and the ugly. J Huntingt Dis. 2016;5(3):217-237. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 9.Quintanilla RA, Johnson GV. Role of mitochondrial dysfunction in the pathogenesis of Huntington's disease. Brain Research Bulletin. 2009;80(4-5):242-247. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 10.Tabrizi S, Cleeter M, Xuereb J, Taanman JW, Cooper J, Schapira A. Biochemical abnormalities and excitotoxicity in Huntington's disease brain. Ann Neurol. 1999;45(1):25-32. [DOI] [PubMed] [Google Scholar]
- 11.Palpagama TH, Waldvogel HJ, Faull RLM, Kwakowsky A. The role of microglia and astrocytes in Huntington's disease. Front Mol Neurosci. 2019;12:258. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 12.Gil JM, Rego AC. Mechanisms of neurodegeneration in Huntington’s disease. Eur J Neurosci. 2008;27(11):2803-2820. [DOI] [PubMed] [Google Scholar]
- 13.Walker FO. Huntington's disease. Lancet. 2007;369(9557):218-228. [DOI] [PubMed] [Google Scholar]
- 14.Roos RAC. Huntington's disease: A clinical review. Orphanet Journal of Rare Diseases. 2010;5:40. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 15.Pringsheim T, Wiltshire K, Day L, Dykeman J, Steeves T, Jette N. The incidence and prevalence of Huntington's disease: A systematic review and meta‐analysis. Mov Disord. 2012;27(9):1083-1091. [DOI] [PubMed] [Google Scholar]
- 16.Warby SC, Montpetit A, Hayden AR, Carroll JB, Butland SL, Visscher H, et al. CAG expansion in the Huntington disease gene is associated with a specific and targetable predisposing haplogroup. Am J Hum Genet. 2009;84(3):351-366. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 17.Mestre T, Ferreira J, Coelho MM, Rosa M, Sampaio C. Therapeutic interventions for symptomatic treatment in Huntington's disease. Cochrane Database Syst Rev. 2009(3). [DOI] [PubMed] [Google Scholar]
- 18.Videnovic A. Treatment of Huntington disease. Curr Treat Options Neurol. 2013;15(4):424-438. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 19.Frank S. Treatment of Huntington’s disease. Neurotherapeutics. 2014;11(1):153-160. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 20.Coppen EM, Roos RA. Current pharmacological approaches to reduce chorea in Huntington’s disease. Drugs. 2017;77(1):29-46. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 21.Ghosh R, Tabrizi SJ. Clinical aspects of Huntington’s disease. In: Behavioral Neurobiology of Huntington's Disease and Parkinson's Disease. Springer; 2013:3-31. [DOI] [PubMed] [Google Scholar]
- 22.Craufurd D, Snowden J. Neuropsychological and neuropsychiatric aspects of Huntington's disease. Oxf Monogr Med Genet. 2002;45:62-94. [Google Scholar]
- 23.van Duijn E, Craufurd D, Hubers AA, Giltay EJ, Bonelli R, Rickards H, et al. Neuropsychiatric symptoms in a European Huntington's disease cohort (REGISTRY). J Neurol Neurosurg Psychiatry. 2014;85(12):1411-1418. [DOI] [PubMed] [Google Scholar]
- 24.Paulsen JS, Nehl C, Hoth KF, Kanz JE, Benjamin M, Conybeare R, et al. Depression and stages of Huntington’s disease. J Neuropsychiatry Clin Neurosci. 2005;17(4):496-502. [DOI] [PubMed] [Google Scholar]
- 25.Tabrizi SJ, Ghosh R, Leavitt BR. Huntingtin lowering strategies for disease modification in Huntington’s disease. Neuron. 2019;101(5):801-819. [DOI] [PubMed] [Google Scholar]
- 26.Evers MM, Konstantinova P. AAV5-miHTT gene therapy for Huntington disease: Lowering both huntingtins. Expet Opin Biol Ther. 2020;20(10):1121-1124. [DOI] [PubMed] [Google Scholar]
- 27.Rodrigues FB, Wild EJ. Huntington’s disease clinical trials corner: April 2020. J Huntingt Dis. 2020;9:185-197. [DOI] [PubMed] [Google Scholar]
- 28.Hoffmann-La Roche . A Study to Evaluate the Efficacy and Safety of Intrathecally Administered RO7234292. (RG6042) in Patients With Manifest Huntington's Disease 2018 [cited 2021 08/01/2021]. Available from: https://clinicaltrials.gov/ct2/show/NCT03761849. [Google Scholar]
- 29.Roche provides update on tominersen programme in manifest Huntington’s disease [press release]. Basel, Switzerland, 22 March 2021 2021.
- 30.Mestre TA, Sampaio C. Huntington disease: Linking pathogenesis to the development of experimental therapeutics. Curr Neurol Neurosci Rep. 2017;17(2):18. [DOI] [PubMed] [Google Scholar]
- 31.Wild EJ, Tabrizi SJ. Therapies targeting DNA and RNA in Huntington's disease. Lancet Neurol. 2017;16(10):837-847. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 32.Denis HL, David LS, Cicchetti F. Antibody-based therapies for Huntington's disease: Current status and future directions. Neurobiol Dis. 2019;132:104569. [DOI] [PubMed] [Google Scholar]
- 33.Maucksch C, Vazey EM, Gordon RJ, Connor B. Stem cell-based therapy for Huntington's disease. J Cell Biochem. 2013;114(4):754-763. [DOI] [PubMed] [Google Scholar]
- 34.Tabrizi SJ, Flower MD, Ross CA, Wild EJ. Huntington disease: New insights into molecular pathogenesis and therapeutic opportunities. Nat Rev Neurol. 2020;16(10):529-546. [DOI] [PubMed] [Google Scholar]
- 35.Kim A, Lalonde K, Truesdell A, Gomes Welter P, Brocardo PS, Rosenstock TR, et al. New avenues for the treatment of Huntington's disease. Int J Mol Sci. 2021;22(16). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 36.Smith Y, Bevan MD, Shink E, Bolam JP. Microcircuitry of the direct and indirect pathways of the basal ganglia. Neuroscience. 1998;86(2):353-387. [DOI] [PubMed] [Google Scholar]
- 37.Albin RL, Reiner A, Anderson KD, Dure LS, Handelin B, Balfour R, et al. Preferential loss of striato-external pallidal projection neurons in presymptomatic Huntington's disease. Ann Neurol. 1992;31(4):425-430. [DOI] [PubMed] [Google Scholar]
- 38.Penney JB, Jr., Young AB. Striatal inhomogeneities and basal ganglia function. Mov Disord. 1986;1(1):3-15. [DOI] [PubMed] [Google Scholar]
- 39.Reiner A, Albin RL, Anderson KD, D'Amato CJ, Penney JB, Young AB. Differential loss of striatal projection neurons in Huntington disease. Proc Natl Acad Sci U S A. 1988;85(15):5733-5737. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 40.Kirkwood SC, Su JL, Conneally PM, Foroud T. Progression of symptoms in the early and middle stages of Huntington disease. Arch Neurol. 2001;58(2):273-278. [DOI] [PubMed] [Google Scholar]
- 41.Hayden MR, Leavitt BR, Yasothan U, Kirkpatrick P. Tetrabenazine. Nature Publishing Group; 2009. [DOI] [PubMed] [Google Scholar]
- 42.Zheng G, Dwoskin LP, Crooks PA. Vesicular monoamine transporter 2: Role as a novel target for drug development. AAPS J. 2006;8(4):E682-E92. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 43.Group HS. Tetrabenazine as antichorea therapy in Huntington disease: A randomized controlled trial. Neurology. 2006;66(3):366-372. [DOI] [PubMed] [Google Scholar]
- 44.Kieburtz K, Penney JB, Corno P, Ranen N, Shoulson I, Feigin A, et al. Unified Huntington’s disease rating scale: Reliability and consistency. Neurology. 2001;11(2):136-142. [Google Scholar]
- 45.Jankovic J, Beach J. Long-term effects of tetrabenazine in hyperkinetic movement disorders. Neurology. 1997;48(2):358-362. [DOI] [PubMed] [Google Scholar]
- 46.Claassen DO, Carroll B, De Boer LM, Wu E, Ayyagari R, Gandhi S, et al. Indirect tolerability comparison of Deutetrabenazine and Tetrabenazine for Huntington disease. J Clinic Movement Disorder. 2017;4(1):3. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 47.Frank S, Testa CM, Stamler D, Kayson E, Davis C, Edmondson MC, et al. Effect of deutetrabenazine on chorea among patients with Huntington disease: A randomized clinical trial. JAMA. 2016;316(1):40-50. [DOI] [PubMed] [Google Scholar]
- 48.Yero T, Rey JA. Tetrabenazine (Xenazine), an FDA-approved treatment option for Huntington’s disease–related chorea. Pharma Therapeutic. 2008;33(12):690. [PMC free article] [PubMed] [Google Scholar]
- 49.Claassen DO, Carroll B, De Boer LM, Wu E, Ayyagari R, Gandhi S, et al. Indirect tolerability comparison of Deutetrabenazine and Tetrabenazine for Huntington disease. J Clin Mov Disord. 2017;4:3. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 50.Schultz JL, Killoran A, Nopoulos PC, Chabal CC, Moser DJ, Kamholz JA. Evaluating depression and suicidality in tetrabenazine users with Huntington disease. Neurology. 2018;91(3):e202-e7. [DOI] [PubMed] [Google Scholar]
- 51.Dorsey E, Brocht AF, Nichols PE, Darwin KC, Anderson KE, Beck CA, et al. Depressed mood and suicidality in individuals exposed to tetrabenazine in a large Huntington disease observational study. J Huntingt Dis. 2013;2(4):509-515. [DOI] [PubMed] [Google Scholar]
- 52.Walker FO. Huntington's disease. Lancet. 2007;369(9557):218-228. [DOI] [PubMed] [Google Scholar]
- 53.Bashir H, Jankovic J. Treatment options for chorea. Expert Review of Neurotherapeutics. 2018;18(1):51-63. [DOI] [PubMed] [Google Scholar]
- 54.Fulton B, Goa KL. Olanzapine. Drugs. 1997;53(2):281-298. [DOI] [PubMed] [Google Scholar]
- 55.Janssen P, Niemegeers C, Awouters F, Schellekens K, Megens A, Meert T. Pharmacology of risperidone (R 64 766), a new antipsychotic with serotonin-S2 and dopamine-D2 antagonistic properties. J Pharmacol Exp Therapeut. 1988;244(2):685-693. [PubMed] [Google Scholar]
- 56.Armstrong MJ, Miyasaki JM. Evidence-based guideline: Pharmacologic treatment of chorea in Huntington disease: Report of the guideline development subcommittee of the American Academy of Neurology. Neurology. 2012;79(6):597-603. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 57.Vijayakumar D, Jankovic J. Drug-induced dyskinesia, part 1: Treatment of levodopa-induced dyskinesia. Drugs. 2016;76(7):759-777. [DOI] [PubMed] [Google Scholar]
- 58.Metman LV, Morris M, Farmer C, Gillespie M, Mosby K, Wuu J, et al. Huntington’s disease: A randomized, controlled trial using the NMDA-antagonist amantadine. Neurology. 2002;59(5):694-699. [DOI] [PubMed] [Google Scholar]
- 59.Rosas HD, Koroshetz WJ, Jenkins BG, Chen YI, Hayden DL, Beal MF, et al. Riluzole therapy in Huntington's disease (HD). Movement disorders. Official Journal of the Movement Disorder Society. 1999;14(2):326-330. [DOI] [PubMed] [Google Scholar]
- 60.Seppi K, Mueller J, Bodner T, Brandauer E, Benke T, Weirich-Schwaiger H, et al. Riluzole in Huntington's disease (HD): An open label study with one year follow up. J Neurol. 2001;248(10):866-869. [DOI] [PubMed] [Google Scholar]
- 61.Duggan L, Fenton M, Rathbone J, Dardennes R, El‐Dosoky A, Indran S. Olanzapine for schizophrenia. Cochrane Database Syst Rev. 2005(2). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 62.Squitieri F, Cannella M, Porcellini A, Brusa L, Simonelli M, Ruggieri S. Short-term effects of olanzapine in Huntington disease. Cognit Behav Neurol. 2001;14(1):69-72. [PubMed] [Google Scholar]
- 63.Eder U, Mangweth B, Ebenbichler C, Weiss E, Hofer A, Hummer M, et al. Association of olanzapine-induced weight gain with an increase in body fat. Am J Psychiatr. 2001;158(10):1719-1722. [DOI] [PubMed] [Google Scholar]
- 64.Goudie AJ, Smith JA, Halford JC. Characterization of olanzapine-induced weight gain in rats. J Psychopharmacol. 2002;16(4):291-296. [DOI] [PubMed] [Google Scholar]
- 65.Ball MP, Coons VB, Buchanan RW. A program for treating olanzapine-related weight gain. Psychiatr Serv. 2001;52(7):967-969. [DOI] [PubMed] [Google Scholar]
- 66.Marder SR, Meibach RC. Risperidone in the treatment of schizophrenia. Am J Psychiatr. 1994. [DOI] [PubMed] [Google Scholar]
- 67.Dazzan P, Morgan KD, Orr K, Hutchinson G, Chitnis X, Suckling J, et al. Different effects of typical and atypical antipsychotics on grey matter in first episode psychosis: The AESOP study. Neuropsychopharmacology. 2005;30(4):765-774. [DOI] [PubMed] [Google Scholar]
- 68.Komossa K, Rummel‐Kluge C, Schwarz S, Schmid F, Hunger H, Kissling W, et al. Risperidone versus other atypical antipsychotics for schizophrenia. Cochrane Database Syst Rev. 2011(1). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 69.Duff K, Beglinger LJ, O'Rourke ME, Nopoulos P, Paulson HL, Paulsen JS. Risperidone and the treatment of psychiatric, motor, and cognitive symptoms in Huntington's disease. Ann Clin Psychiatr. 2008;20(1):1-3. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 70.Conley RR, Mahmoud R. A randomized double-blind study of risperidone and olanzapine in the treatment of schizophrenia or schizoaffective disorder. Am J Psychiatr. 2001;158(5):765-774. [DOI] [PubMed] [Google Scholar]
- 71.Naarding P, Kremer H, Zitman F. Huntington’s disease: A review of the literature on prevalence and treatment of neuropsychiatric phenomena. Eur Psychiatr. 2001;16(8):439-445. [DOI] [PubMed] [Google Scholar]
- 72.Hyttel J. Pharmacological Characterization of Selective Serotonin Reuptake Inhibitors (SSRIs). International clinical psychopharmacology; 1994. [DOI] [PubMed] [Google Scholar]
- 73.Preskorn SH. Clinically relevant pharmacology of selective serotonin reuptake inhibitors. Clin Pharmacokinet. 1997;32(1):1-21. [DOI] [PubMed] [Google Scholar]
- 74.Goodnick PJ, Goldstein BJ. Selective serotonin reuptake inhibitors in affective disorders—I. Basic pharmacology. J Psychopharmacol. 1998;12(4_suppl l):5-S20. [DOI] [PubMed] [Google Scholar]
- 75.Taylor C, Fricker AD, Devi LA, Gomes I. Mechanisms of action of antidepressants: From neurotransmitter systems to signaling pathways. Cell Signal. 2005;17(5):549-557. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 76.Chu A, Wadhwa R. Selective Serotonin Reuptake Inhibitors. In: StatPearls. Treasure Island (FL): StatPearls Publishing. Copyright © 2021, StatPearls Publishing LLC.; 2021. [PubMed]
- 77.Khushboo SB. Antidepressants: Mechanism of action, toxicity and possible amelioration. J Applied Biotechnol Bioengineer. 2017;3(5). [Google Scholar]
- 78.Hirschfeld R. Long-term side effects of SSRIs: Sexual dysfunction and weight gain. J Clin Psychiatr. 2003;64:20-24. [PubMed] [Google Scholar]
- 79.Gibbons RD, Brown CH, Hur K, Marcus SM, Bhaumik DK, Erkens JA, et al. Early evidence on the effects of regulators’ suicidality warnings on SSRI prescriptions and suicide in children and adolescents. Am J Psychiatr. 2007;164(9):1356-1363. [DOI] [PubMed] [Google Scholar]
- 80.Garland EJ, Kutcher S, Virani A, Elbe D. Update on the use of SSRIs and SNRIs with children and adolescents in clinical practice. J Canad Acad Child Adol Psychiat. 2016;25(1):4. [PMC free article] [PubMed] [Google Scholar]
- 81.Moulton CD, Hopkins C, Bevan‐Jones WR. Systematic review of pharmacological treatments for depressive symptoms in Huntington's disease. Mov Disord. 2014;29(12):1556-1561. [DOI] [PubMed] [Google Scholar]
- 82.Beglinger LJ, Adams WH, Langbehn D, Fiedorowicz JG, Jorge R, Biglan K, et al. Results of the citalopram to enhance cognition in Huntington disease trial. Mov Disord. 2014;29(3):401-405. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 83.De Marchi N, Daniele F, Ragone MA. Fluoxetine in the treatment of Huntington's disease. Psychopharmacology. 2001;153(2):264-266. [DOI] [PubMed] [Google Scholar]
- 84.Benfield P, Heel RC, Lewis SP. Fluoxetine. Drugs. 1986;32(6):481-508. [DOI] [PubMed] [Google Scholar]
- 85.Ranen NG, Lipsey JR, Treisman G, Ross CA. Sertraline in the treatment of severe aggressiveness in Huntington's disease. J Neuropsychiatry Clin Neurosci. 1996;8(3):338-340. [DOI] [PubMed] [Google Scholar]
- 86.Patzold T, Brüne M. Obsessive compulsive disorder in Huntington disease: A case of isolated obsessions successfully treated with sertraline. Cognit Behav Neurol. 2002;15(3):216-219. [PubMed] [Google Scholar]
- 87.Monroe RR. Anticonvulsants in the treatment of aggression. J Nerv Ment Dis. 1975;160:119-126. [DOI] [PubMed] [Google Scholar]
- 88.Stanford MS, Helfritz LE, Conklin SM, Villemarette-Pittman NR, Greve KW, Adams D, et al. A comparison of anticonvulsants in the treatment of impulsive aggression. Exp Clin Psychopharmacol. 2005;13(1):72. [DOI] [PubMed] [Google Scholar]
- 89.Grunze HC. Anticonvulsants in bipolar disorder. J Ment Health. 2010;19(2):127-141. [DOI] [PubMed] [Google Scholar]
- 90.Yatham LN. Newer anticonvulsants in the treatment of bipolar disorder. J Clin Psychiatr. 2004;65:28-35. [PubMed] [Google Scholar]
- 91.Bruno A, Micò U, Pandolfo G, Mallamace D, Abenavoli E, Di Nardo F, et al. Lamotrigine augmentation of serotonin reuptake inhibitors in treatment-resistant obsessive–compulsive disorder: A double-blind, placebo-controlled study. J Psychopharmacol. 2012;26(11):1456-1462. [DOI] [PubMed] [Google Scholar]
- 92.Wang HR, Woo YS, Bahk WM. Potential role of anticonvulsants in the treatment of obsessive–compulsive and related disorders. Psychiatr Clin Neurosci. 2014;68(10):723-732. [DOI] [PubMed] [Google Scholar]
- 93.Macdonald R, McLean M. Anticonvulsant drugs: Mechanisms of action. Adv Neurol. 1986;44:713. [PubMed] [Google Scholar]
- 94.Appleton RE, Freeman A, Cross JH. Diagnosis and management of the epilepsies in children: A summary of the partial update of the 2012 NICE epilepsy guideline. Arch Dis Child. 2012;97(12):1073-1076. [DOI] [PubMed] [Google Scholar]
- 95.Hammer GD, McPhee SJ. Pathophysiology of Disease: An Introduction to Clinical Medicine. Univerza v Ljubljani, Medicinska fakulteta; 2014. [Google Scholar]
- 96.Stefan H, Feuerstein T. Novel anticonvulsant drugs. Pharmacol Ther. 2007;113(1):165-183. [DOI] [PubMed] [Google Scholar]
- 97.Grunze HCR. The effectiveness of anticonvulsants in psychiatric disorders. Dialogues Clin Neurosci. 2008;10(1):77-89. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 98.Gould TD, Chen G, Manji HK. Mood stabilizer psychopharmacology. Clin Neurosci Res. 2002;2(3-4):193-212. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 99.Ross CA, Tabrizi SJ. Huntington's disease: From molecular pathogenesis to clinical treatment. Lancet Neurol. 2011;10(1):83-98. [DOI] [PubMed] [Google Scholar]
- 100.Chiu C-T, Liu G, Leeds P, Chuang D-M. Combined treatment with the mood stabilizers lithium and valproate produces multiple beneficial effects in transgenic mouse models of Huntington's disease. Neuropsychopharmacology. 2011;36(12):2406-2421. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 101.Scheuing L, Chiu C-T, Liao H-M, Linares GR, Chuang D-M. Preclinical and clinical investigations of mood stabilizers for Huntington's disease: What have we learned? Int J Biol Sci. 2014;10(9):1024. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 102.Lees G, Leach MJ. Studies on the mechanism of action of the novel anticonvulsant lamotrigine (Lamictal) using primary neuroglial cultures from rat cortex. Brain Res. 1993;612(1-2):190-199. [DOI] [PubMed] [Google Scholar]
- 103.Wu J, Tang T, Bezprozvanny I. Evaluation of clinically relevant glutamate pathway inhibitors in in vitro model of Huntington's disease. Neurosci Lett. 2006;407(3):219-223. [DOI] [PubMed] [Google Scholar]
- 104.Kremer B, Clark C, Almqvist E, Raymond L, Graf P, Jacova C, et al. Influence of lamotrigine on progression of early Huntington disease: A randomized clinical trial. Neurology. 1999;53(5):1000. [DOI] [PubMed] [Google Scholar]
- 105.Ambrósio AF, Soares-da-Silva P, Carvalho CM, Carvalho AP. Mechanisms of action of carbamazepine and its derivatives, oxcarbazepine, BIA 2-093, and BIA 2-024. Neurochem Res. 2002;27(1-2):121-130. [DOI] [PubMed] [Google Scholar]
- 106.Lang DG, Wang CM, Cooper B. Lamotrigine, phenytoin and carbamazepine interactions on the sodium current present in N4TG1 mouse neuroblastoma cells. J Pharmacol Exp Therapeut. 1993;266(2):829-835. [PubMed] [Google Scholar]
- 107.Devi K, George S, Criton S, Suja V, Sridevi P. Carbamazepine-The commonest cause of toxic epidermal necrolysis and Stevens-Johnson syndrome: A study of 7 years. Indian J Dermatol Venereol Leprol. 2005;71(5):325. [DOI] [PubMed] [Google Scholar]
- 108.Bilney B, Morris ME, Perry A. Effectiveness of physiotherapy, occupational therapy, and speech pathology for people with Huntington's disease: A systematic review. Neurorehabilitation Neural Repair. 2003;17(1):12-24. [DOI] [PubMed] [Google Scholar]
- 109.Mirek E, Filip M, Banaszkiewicz K, Rudzińska M, Szymura J, Pasiut S, et al. The effects of physiotherapy with PNF concept on gait and balance of patients with Huntington's disease–pilot study. Neurol Neurochir Pol. 2015;49(6):354-357. [DOI] [PubMed] [Google Scholar]
- 110.Quinn L, Busse M. Physiotherapy clinical guidelines for Huntington’s disease. Neurodegener Dis Manag. 2012;2(1):21-31. [Google Scholar]
- 111.Busse ME, Khalil H, Quinn L, Rosser AE. Physical therapy intervention for people with Huntington disease. Phys Ther. 2008;88(7):820-831. [DOI] [PubMed] [Google Scholar]
- 112.Barkmeier-Kraemer JM, Clark HM. Speech-language pathology evaluation and management of hyperkinetic disorders affecting speech and swallowing function. Tremor Hyperkinetic Movement. 2017;7:489. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 113.Butnaru H, Krueger WO. Communication Device and Method for Deaf and Mute Persons. Google Patents; 2001. [Google Scholar]
- 114.Pennington L, Goldbart J, Marshall J. Speech and language therapy to improve the communication skills of children with cerebral palsy. Cochrane Database Syst Rev. 2004(2). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 115.Żukiewicz-Sobczak W, Król R, Wróblewska P, Piątek J, Gibas-Dorna M. Huntington disease–principles and practice of nutritional management. Neurol Neurochir Pol. 2014;48(6):442-448. [DOI] [PubMed] [Google Scholar]
- 116.Bachoud-Lévi A-C, Ferreira J, Massart R, Youssov K, Rosser A, Busse M, et al. International Guidelines for the Treatment of Huntington's Disease. Front Neurol. 2019;10(710). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 117.Nagel SJ, Machado AG, Gale JT, Lobel DA, Pandya M. Preserving cortico-striatal function: Deep brain stimulation in Huntington’s disease. Front Syst Neurosci. 2015;9(32). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 118.Sharma M, Deogaonkar M. Deep brain stimulation in Huntington’s disease: Assessment of potential targets. J Clin Neurosci. 2015;22(5):812-817. [DOI] [PubMed] [Google Scholar]
- 119.Temel Y, Cao C, Vlamings R, Blokland A, Ozen H, Steinbusch HW, et al. Motor and cognitive improvement by deep brain stimulation in a transgenic rat model of Huntington's disease. Neurosci Lett. 2006;406(1-2):138-141. [DOI] [PubMed] [Google Scholar]
- 120.Wojtecki L, Groiss SJ, Hartmann CJ, Elben S, Omlor S, Schnitzler A, et al. Deep brain stimulation in Huntington’s disease—preliminary evidence on pathophysiology, efficacy and safety. Brain Sciences. 2016;6(3):38. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 121.Gonzalez V, Cif L, Biolsi B, Garcia-Ptacek S, Seychelles A, Sanrey E, et al. Deep brain stimulation for Huntington's disease: Long-term results of a prospective open-label study. J Neurosurg. 2014;121(1):114-122. [DOI] [PubMed] [Google Scholar]
- 122.Fenoy AJ, Simpson RK. Risks of common complications in deep brain stimulation surgery: Management and avoidance. J Neurosurg. 2014;120(1):132-139. [DOI] [PubMed] [Google Scholar]
- 123.Ghosh R, Tabrizi SJ. Clinical aspects of Huntington's disease. Curr Top Behav Neurosci. 2015;22:3-31. [DOI] [PubMed] [Google Scholar]
- 124.Scoles DR, Pulst SM. Oligonucleotide therapeutics in neurodegenerative diseases. RNA Biol. 2018;15(6):707-714. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 125.Rodrigues FB, Wild EJ. Huntington’s disease clinical trials corner: February 2018. J Huntingt Dis. 2018;7(1):89-98. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 126.Marxreiter F, Stemick J, Kohl Z. Huntingtin lowering strategies. Int J Mol Sci. 2020;21(6):2146. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 127.Kordasiewicz HB, Stanek LM, Wancewicz EV, Mazur C, McAlonis MM, Pytel KA, et al. Sustained therapeutic reversal of Huntington's disease by transient repression of huntingtin synthesis. Neuron. 2012;74(6):1031-1044. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 128.Ionis Pharmaceuticals I . Safety, Tolerability, Pharmacokinetics, and Pharmacodynamics of ISIS 443139 in Participants with Early Manifest Huntington's Disease. National Institutes of Health; 2019. [updated May 31 2019; cited 2021 08/01/2021]. Available from: https://clinicaltrials.gov/ct2/show/NCT02519036. [Google Scholar]
- 129.Tabrizi SJ, Leavitt BR, Landwehrmeyer GB, Wild EJ, Saft C, Barker RA, et al. Targeting Huntingtin expression in patients with Huntington's disease. N Engl J Med. 2019;380(24):2307-2316. [DOI] [PubMed] [Google Scholar]
- 130.Dhuri K, Bechtold C, Quijano E, Pham H, Gupta A, Vikram A, et al. Antisense oligonucleotides: An emerging area in drug discovery and development. J Clin Med. 2020;9(6):2004. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 131.Rodrigues FB, Ferreira JJ, Wild EJ. Huntington's disease clinical trials corner: June 2019. J Huntingt Dis. 2019;8(3):363-371. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 132.Hoffmann-La Roche . A Study to Evaluate the Safety, Tolerability, Pharmacokinetics, and Pharmacodynamics of RO7234292 (ISIS 443139) in Huntington's Disease Patients Who Participated in Prior Investigational Studies of RO7234292 (ISIS 443139) 2019 [updated June 9 20202; cited 2021 08/01/2021]. Available from. https://clinicaltrials.gov/ct2/show/NCT03342053.
- 133.Rodrigues FB, Quinn L, Wild EJ. Huntington’s disease clinical trials corner: January 2019. J Huntingt Dis. 2019;8:115-125. [DOI] [PubMed] [Google Scholar]
- 134.Hoffmann-La Roche . An Open-Label Extension Study to Evaluate Long-Term Safety and Tolerability of RO7234292 (RG6042) in Huntington's Disease Patients Who Participated in Prior Roche and Genentech Sponsored Studies 2019 [updated December 10 2020; cited 2021 08/01/2021]. Available from. https://clinicaltrials.gov/ct2/show/NCT03842969.
- 135.Hoffmann-La Roche . A Study to Investigate the Pharmacokinetics and Pharmacodynamics of RO7234292 (RG6042) in CSF and Plasma, and Safety and Tolerability Following Intrathecal Administration in Patients With Huntington's Disease 2019 [cited 2021 08/01/2021]. Available from. https://clinicaltrials.gov/ct2/show/NCT04000594.
- 136.Hersch S, Claassen D, Edmondson M, Wild E, Guerciolini R, Panzara M. Multicenter, randomized, double-blind, placebo-controlled phase 1b/2a studies of WVE-120101 and WVE-120102 in patients with Huntington’s disease (P2.006). Neurology. 2017;88(16 suppl):P2.006. [Google Scholar]
- 137.Kaemmerer WF, Grondin RC. The effects of huntingtin-lowering: What do we know so far? Degener Neurol Neuromuscul Dis. 2019;9:3-17. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 138.Wave Life Sciences Announces Topline Data and Addition of Higher Dose Cohort in Ongoing Phase 1b/2a PRECISION-HD2 Trial in Huntington’s Disease [press release]. Cambridge, Massachusetts: Wave Life Sciences, December 30 2019 2019.
- 139.Wave Life Sciences Provides Update on Phase 1b/2a PRECISION-HD Trials. [press release]. March 29 2021.
- 140.Wave Life Sciences Reports Third Quarter 2020 Financial Results and Provides Business Update [press release]. Cambridge, Massachusetts: Wave Life Sciences, November 9 2020 2020.
- 141.Dale E, Frank-Kamenetsky M, Taborn K, Aklilu K, Liu Y, Berkovitch S, et al. Stereopure oligonucleotides for the selective silencing of mutant Huntingtin (4703). Neurology. 2020;94(15 suppl):4703. [Google Scholar]
- 142.Evers MM, Pepers BA, van Deutekom JCT, Mulders SAM, den Dunnen JT, Aartsma-Rus A, et al. Targeting several CAG expansion diseases by a single antisense oligonucleotide. PLoS One. 2011;6(9):e24308. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 143.Datson NA, González-Barriga A, Kourkouta E, Weij R, van de Giessen J, Mulders S, et al. The expanded CAG repeat in the huntingtin gene as target for therapeutic RNA modulation throughout the HD mouse brain. PLoS One. 2017;12(2):e0171127. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 144.Li JY, Popovic N, Brundin P. The use of the R6 transgenic mouse models of Huntington's disease in attempts to develop novel therapeutic strategies. NeuroRx. 2005;2(3):447-464. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 145.Menalled LB, Kudwa AE, Miller S, Fitzpatrick J, Watson-Johnson J, Keating N, et al. comprehensive behavioral and molecular characterization of a new knock-in mouse model of Huntington’s disease: zQ175. PLoS One. 2012;7(12):e49838. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 146.Skotte NH, Southwell AL, Østergaard ME, Carroll JB, Warby SC, Doty CN, et al. Allele-specific suppression of mutant huntingtin using antisense oligonucleotides: Providing a therapeutic option for all Huntington disease patients. PLoS One. 2014;9(9):e107434. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 147.Southwell AL, Skotte NH, Kordasiewicz HB, Østergaard ME, Watt AT, Carroll JB, et al. In vivo evaluation of candidate allele-specific mutant huntingtin gene silencing antisense oligonucleotides. Mol Ther. 2014;22(12):2093-2106. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 148.Southwell AL, Kordasiewicz HB, Langbehn D, Skotte NH, Parsons MP, Villanueva EB, et al. Huntingtin suppression restores cognitive function in a mouse model of Huntington’s disease. Sci Transl Med. 2018;10(461):eaar3959. [DOI] [PubMed] [Google Scholar]
- 149.Evers MM, Tran H-D, Zalachoras I, Meijer OC, den Dunnen JT, van Ommen G-JB, et al. Preventing formation of toxic N-terminal Huntingtin fragments through antisense oligonucleotide-mediated protein modification. Nucleic Acid Therapeut. 2014;24(1):4-12. [DOI] [PubMed] [Google Scholar]
- 150.Triplet Therapeutics Presents Preclinical Data for TTX-3360, Announces Equity Pledge at the CHDI Foundation’s 16th Annual HD Therapeutics Conference [press release]. April 27 2021.
- 151.McBride JL, Boudreau RL, Harper SQ, Staber PD, Monteys AM, Martins I, et al. Artificial miRNAs mitigate shRNA-mediated toxicity in the brain: Implications for the therapeutic development of RNAi. Proc Natl Acad Sci USA. 2008;105(15):5868-5873. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 152.Grimm D, Streetz KL, Jopling CL, Storm TA, Pandey K, Davis CR, et al. Fatality in mice due to oversaturation of cellular microRNA/short hairpin RNA pathways. Nature. 2006;441(7092):537-541. [DOI] [PubMed] [Google Scholar]
- 153.Aguiar S, van der Gaag B, Cortese FAB. RNAi mechanisms in Huntington's disease therapy: siRNA versus shRNA. Transl Neurodegener. 2017;6:30. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 154.Miniarikova J, Zanella I, Huseinovic A, van der Zon T, Hanemaaijer E, Martier R, et al. Design, characterization, and lead selection of therapeutic miRNAs targeting Huntingtin for development of gene therapy for Huntington's disease. Mol Ther Nucleic Acids. 2016;5(3):e297-e. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 155.Miniarikova J, Zimmer V, Martier R, Brouwers CC, Pythoud C, Richetin K, et al. AAV5-miHTT gene therapy demonstrates suppression of mutant Huntingtin aggregation and neuronal dysfunction in a rat model of Huntington's disease. Gene Therapy. 2017;24(10):630-639. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 156.Keskin S, Brouwers CC, Sogorb-Gonzalez M, Martier R, Depla JA, Vallès A, et al. AAV5-miHTT lowers Huntingtin mRNA and protein without Off-target effects in patient-derived neuronal cultures and astrocytes. Mol Ther Method Clinic Dev. 2019;15:275-284. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 157.Southwell AL, Skotte NH, Villanueva EB, Østergaard ME, Gu X, Kordasiewicz HB, et al. A novel humanized mouse model of Huntington disease for preclinical development of therapeutics targeting mutant huntingtin alleles. Hum Mol Genet. 2017;26(6):1115-1132. [DOI] [PubMed] [Google Scholar]
- 158.Caron NS, Southwell AL, Brouwers CC, Cengio LD, Xie Y, Black HF, et al. Potent and sustained huntingtin lowering via AAV5 encoding miRNA preserves striatal volume and cognitive function in a humanized mouse model of Huntington disease. Nucleic Acids Res. 2020;48(1):36-54. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 159.Spronck EA, Brouwers CC, Vallès A, de Haan M, Petry H, van Deventer SJ, et al. AAV5-miHTT gene therapy demonstrates sustained Huntingtin lowering and functional improvement in Huntington disease mouse models. Mol Therapy Method Clinic Dev. 2019;13:334-343. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 160.Evers MM, Miniarikova J, Juhas S, Vallès A, Bohuslavova B, Juhasova J, et al. AAV5-miHTT gene therapy demonstrates broad distribution and strong human mutant Huntingtin lowering in a Huntington's disease minipig model. Mol Ther. 2018;26(9):2163-2177. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 161.Spronck EA, Vallès A, Lampen MH, Montenegro-Miranda PS, Keskin S, Heijink L, et al. Intrastriatal administration of AAV5-miHTT in non-human primates and rats Is well tolerated and results in miHTT transgene expression in key areas of Huntington disease pathology. Brain Sci. 2021;11(2):129. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 162.UniQure. Safety and Proof-of-Concept (POC) Study With AMT-130 in Adults With Early Manifest Huntington Disease 2019 [updated October 28 2020; cited 2021 09/01/2021]. Available from: https://clinicaltrials.gov/ct2/show/NCT04120493.
- 163.uniQure Announces Completion of Enrollment in First Cohort of Phase I/II Clinical Trial of AMT-130 for the Treatment of Huntington’s Disease [press release]. Amsterdam, The Netherlands, April 5, 2021 2021.
- 164.UniQure Announces Positive Recommendation from Data Safety Monitoring Board of Phase I/II Clinical Trial of AMT-130 for the Treatment of Huntington’s Disease [press release]. Lexington, Massachusetts; Amsterdam, The Netherlands,: UniQure, February 8 2021 2021.
- 165.Stanek LM, Sardi SP, Mastis B, Richards AR, Treleaven CM, Taksir T, et al. Silencing mutant huntingtin by adeno-associated virus-mediated RNA interference ameliorates disease manifestations in the YAC128 mouse model of Huntington's disease. Hum Gene Ther. 2014;25(5):461-474. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 166.Ghilan M, Bostrom CA, Hryciw BN, Simpson JM, Christie BR, Gil-Mohapel J. YAC128 Huntington's disease transgenic mice show enhanced short-term hippocampal synaptic plasticity early in the course of the disease. Brain Research. 2014;1581:117-128. [DOI] [PubMed] [Google Scholar]
- 167.Voyager Therapeutics Announces Preclinical Data For Huntington’s Disease And Amyotrophic Lateral Sclerosis Programs At The Congress Of The European Society Of Gene And Cell Therapy [press release]. Cambridge, Massachusetts, October 16 2018.
- 168.Voyager Therapeutics Provides Regulatory Update On VY-HTT01 Program [press release]. Cambridge, Massachusetts: Voyager Therapeutics, October 13 2020.
- 169.Harper SQ, Staber PD, He X, Eliason SL, Martins IH, Mao Q, et al. RNA interference improves motor and neuropathological abnormalities in a Huntington's disease mouse model. Proc Natl Acad Sci USA. 2005;102(16):5820-5825. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 170.Franich NR, Fitzsimons HL, Fong DM, Klugmann M, During MJ, Young D. AAV vector-mediated RNAi of mutant huntingtin expression is neuroprotective in a novel genetic rat model of Huntington's disease. Mol Ther. 2008;16(5):947-956. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 171.McBride JL, Pitzer MR, Boudreau RL, Dufour B, Hobbs T, Ojeda SR, et al. Preclinical safety of RNAi-mediated HTT suppression in the rhesus macaque as a potential therapy for Huntington's disease. Mol Ther. 2011;19(12):2152-2162. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 172.Bhattacharyya A, ed. Identification and development of orally administered, CNS-penetrant small molecules that lower huntingtin protein levels by inducing a novel splicing event that alters the stability of huntingtin mRNA. CHDI HD Therapeutics Conference; 2019; Palm Springs, California.
- 173.PTC Therapeutics Announces that PTC518 Has Entered into a Phase 1 Clinical Trial for the Huntington's Disease Program [press release]. South Plainfield, New Jersey2020.
- 174.Gray M, Shirasaki DI, Cepeda C, André VM, Wilburn B, Lu X-H, et al. Full-length human mutant huntingtin with a stable polyglutamine repeat can elicit progressive and selective neuropathogenesis in BACHD mice. J Neurosci. 2008;28(24):6182-6195. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 175.Therapeutics PTC. PTC518 Phase 1 Huntington's Disease Program Update 2021 [cited 2022 January]. Available from:. https://ir.ptcbio.com/events/event-details/ptc518-phase-1-huntingtons-disease-program-update.
- 176.Shorrock HK, Gillingwater TH, Groen EJN. Overview of current drugs and molecules in development for spinal muscular atrophy therapy. Drugs. 2018;78(3):293-305. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 177.Novartis receives US Food and Drug Administration (FDA) Orphan Drug Designation for branaplam (LMI070) in Huntington’s disease (HD) [press release]. Basel, Switzerland, October 21 2020.
- 178.Liu CR, Chang CR, Chern Y, Wang TH, Hsieh WC, Shen WC, et al. Spt4 is selectively required for transcription of extended trinucleotide repeats. Cell. 2012;148(4):690-701. [DOI] [PubMed] [Google Scholar]
- 179.Cheng H-M, Chern Y, Chen IH, Liu C-R, Li S-H, Chun SJ, et al. Effects on murine behavior and lifespan of selectively decreasing expression of mutant huntingtin allele by supt4h knockdown. PLoS Genetics. 2015;11(3):e1005043. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 180.Zhou X, Li G, Kaplan A, Gaschler MM, Zhang X, Hou Z, et al. Small molecule modulator of protein disulfide isomerase attenuates mutant huntingtin toxicity and inhibits endoplasmic reticulum stress in a mouse model of Huntington's disease. Hum Mol Genet. 2018;27(9):1545-1555. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 181.Bard J, Wall MD, Lazari O, Arjomand J, Munoz-Sanjuan I. Advances in huntington disease drug discovery: Novel approaches to model disease phenotypes. J Biomol Screen. 2014;19(2):191-204. [DOI] [PubMed] [Google Scholar]
- 182.Khan E, Mishra SK, Mishra R, Mishra A, Kumar A. Discovery of a potent small molecule inhibiting Huntington's disease (HD) pathogenesis via targeting CAG repeats RNA and Poly Q protein. Sci Rep. 2019;9(1):16872. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 183.Khan E, Tawani A, Mishra SK, Verma AK, Upadhyay A, Kumar M, et al. Myricetin reduces toxic level of CAG repeats RNA in Huntington's disease (HD) and spino cerebellar ataxia (SCAs). ACS Chem Biol. 2018;13(1):180-188. [DOI] [PubMed] [Google Scholar]
- 184.Khan E, Biswas S, Mishra SK, Mishra R, Samanta S, Mishra A, et al. Rationally designed small molecules targeting toxic CAG repeat RNA that causes Huntington's disease (HD) and spinocerebellar ataxia (SCAs). Biochimie. 2019;163:21-32. [DOI] [PubMed] [Google Scholar]
- 185.Lee J-M, Zhang J, Su AI, Walker JR, Wiltshire T, Kang K, et al. A novel approach to investigate tissue-specific trinucleotide repeat instability. BMC Syst Biol. 2010;4:29. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 186.Garriga-Canut M, Agustín-Pavón C, Herrmann F, Sánchez A, Dierssen M, Fillat C, et al. Synthetic zinc finger repressors reduce mutant huntingtin expression in the brain of R6/2 mice. Proc Natl Acad Sci USA. 2012;109(45):E3136-E3145. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 187.Agustín-Pavón C, Mielcarek M, Garriga-Canut M, Isalan M. Deimmunization for gene therapy: Host matching of synthetic zinc finger constructs enables long-term mutant Huntingtin repression in mice. Mol Neurodegener. 2016;11(1):64. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 188.Zeitler B, Froelich S, Marlen K, Shivak DA, Yu Q, Li D, et al. Allele-selective transcriptional repression of mutant HTT for the treatment of Huntington’s disease. Nat Med. 2019;25(7):1131-1142. [DOI] [PubMed] [Google Scholar]
- 189.Menalled LB. Knock-in mouse models of Huntington's disease. NeuroRx. 2005;2(3):465-470. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 190.Nemudryi AA, Valetdinova KR, Medvedev SP, Zakian SM. TALEN and CRISPR/Cas genome editing systems: Tools of discovery. Acta naturae. 2014;6(3):19-40. [PMC free article] [PubMed] [Google Scholar]
- 191.Fink KD, Deng P, Gutierrez J, Anderson JS, Torrest A, Komarla A, et al. Allele-specific reduction of the mutant huntingtin allele using transcription activator-like effectors in human Huntington's disease fibroblasts. Cell Transplantation. 2016;25(4):677-686. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 192.Richard G-F, Viterbo D, Khanna V, Mosbach V, Castelain L, Dujon B. Highly specific contractions of a single CAG/CTG trinucleotide repeat by TALEN in yeast. PLoS One. 2014;9(4):e95611. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 193.Dabrowska M, Juzwa W, Krzyzosiak WJ, Olejniczak M. Precise excision of the CAG tract from the Huntingtin gene by Cas9 nickases. Front Neurosci. 2018;12:75. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 194.Xu X, Tay Y, Sim B, Yoon S-I, Huang Y, Ooi J, et al. Reversal of phenotypic abnormalities by CRISPR/Cas9-mediated gene correction in Huntington disease patient-derived induced pluripotent stem cells. Stem Cell Reports. 2017;8(3):619-633. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 195.Shin JW, Kim K-H, Chao MJ, Atwal RS, Gillis T, MacDonald ME, et al. Permanent inactivation of Huntington's disease mutation by personalized allele-specific CRISPR/Cas9. Hum Mol Genet. 2016;25(20):4566-4576. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 196.Yang S, Chang R, Yang H, Zhao T, Hong Y, Kong HE, et al. CRISPR/Cas9-mediated gene editing ameliorates neurotoxicity in mouse model of Huntington's disease. J Clin Invest. 2017;127(7):2719-2724. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 197.Ekman F, Ojala D, Adil M, Lopez P, Schaffer D, Gaj T. CRISPR-Cas9-mediated genome editing increases lifespan and improves motor deficits in a Huntington’s disease mouse model. Mol Ther Nucleic Acids. 2019;17:829-839. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 198.Monteys AM, Ebanks SA, Keiser MS, Davidson BL. CRISPR/Cas9 editing of the mutant Huntingtin Allele In vitro and in vivo. Mol Ther Therapy. 2017;25(1):12-23. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 199.Salado-Manzano C, Perpiña U, Straccia M, Molina-Ruiz FJ, Cozzi E, Rosser AE, et al. Is the immunological response a bottleneck for cell therapy in neurodegenerative diseases? Front Cell Neurosci. 2020;14:250. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 200.Azidus Brasil. Safety Evaluation of Cellavita HD Administered Intravenously in Participants With Huntington's Disease (SAVE-DH) 2016 [cited 2021 13 January]. Available from: https://clinicaltrials.gov/ct2/show/NCT02728115.
- 201.Azidus Brasil. Dose-response Evaluation of the Cellavita HD Product in Patients With Huntington's Disease (ADORE-DH) 2017. Available from: https://clinicaltrials.gov/ct2/show/NCT03252535.
- 202.Azidus Brasil. Clinical Extension Study for Safety and Efficacy Evaluation of Cellavita-HD Administration in Huntington's Patients. (ADORE-EXT) 2020 [cited 2021 13 January]. Available from: https://clinicaltrials.gov/ct2/show/NCT04219241.
- 203.Rosser AE, Barker RA, Harrower T, Watts C, Farrington M, Ho AK, et al. Unilateral transplantation of human primary fetal tissue in four patients with Huntington’s disease: NEST-UK safety report ISRCTN no 36485475. J Neurol Neurosurg Psychiatr. 2002;73(6):678-685. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 204.Barker RA, Mason SL, Harrower TP, Swain RA, Ho AK, Sahakian BJ, et al. The long-term safety and efficacy of bilateral transplantation of human fetal striatal tissue in patients with mild to moderate Huntington's disease. J Neurol Neurosurg Psychiatr. 2013;84(6):657-665. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 205.Bachoud-Levi A-C, Remy P, Nguyen J-P, Brugieres P, Lefaucheur J-P, Bourdet C, et al. Motor and cognitive improvements in patients with Huntington's disease after neural transplantation. Lancet. 2000;356(9246):1975. [DOI] [PubMed] [Google Scholar]
- 206.Paganini M, Biggeri A, Romoli AM, Mechi C, Ghelli E, Berti V, et al. Fetal striatal grafting slows motor and cognitive decline of Huntington's disease. J Neurol Neurosurg Psychiatr. 2014;85(9):974-981. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 207.Drew C, Rosser A, Gray W. A single site, open label, phase I study to assess the safety and feasibility of foetal cell transplants in the striatum of people with Huntington’s disease 2017. Available from:. http://www.isrctn.com/ISRCTN52651778.
- 208.Bachoud-Lévi AC, Bourdet C, Brugières P, Nguyen JP, Grandmougin T, Haddad B, et al. Safety and tolerability assessment of intrastriatal neural allografts in five patients with Huntington's disease. Exp Neurol. 2000;161(1):194-202. [DOI] [PubMed] [Google Scholar]
- 209.Maxan A, Mason S, Saint-Pierre M, Smith E, Ho A, Harrower T, et al. Outcome of cell suspension allografts in a patient with Huntington's disease. Ann Neurol. 2018;84(6):950-956. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 210.Krystkowiak P, Gaura V, Labalette M, Rialland A, Remy P, Peschanski M, et al. Alloimmunisation to donor antigens and immune rejection following foetal neural grafts to the brain in patients with Huntington's disease. PLoS One. 2007;2(1):e166. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 211.Krebs SS, Trippel M, Prokop T, Omer TN, Landwehrmeyer B, Weber WA, et al. Immune response after striatal engraftment of fetal neuronal cells in patients with Huntington’s disease: Consequences for cerebral transplantation programs. Clinic Exp Neuroimmunol. 2011;2(2):25-32. [Google Scholar]
- 212.Bachoud-Lévi AC, Gaura V, Brugières P, Lefaucheur JP, Boissé MF, Maison P, et al. Effect of fetal neural transplants in patients with Huntington's disease 6 years after surgery: A long-term follow-up study. Lancet Neurol. 2006;5(4):303-309. [DOI] [PubMed] [Google Scholar]
- 213.Cisbani G, Freeman TB, Soulet D, Saint-Pierre M, Gagnon D, Parent M, et al. Striatal allografts in patients with Huntington's disease: Impact of diminished astrocytes and vascularization on graft viability. Brain. 2013;136(Pt 2):433-443. [DOI] [PubMed] [Google Scholar]
- 214.Cho IK, Hunter CE, Ye S, Pongos AL, Chan AWS. Combination of stem cell and gene therapy ameliorates symptoms in Huntington’s disease mice. Npj Regenerative Medicine. 2019;4(1):7. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 215.Lee ST, Chu K, Jung KH, Im WS, Park JE, Lim HC, et al. Slowed progression in models of Huntington disease by adipose stem cell transplantation. Ann Neurol. 2009;66(5):671-681. [DOI] [PubMed] [Google Scholar]
- 216.Welter R. Autologous Stem/Stromal Cells in Neurological Disorders and Disease (NDD) 2017 [cited 2021 20 February]. Available from. https://clinicaltrials.gov/ct2/show/NCT03297177.
- 217.Haddad MS, Wenceslau CV, Pompeia C, Kerkis I. Cell-based technologies for Huntington's disease. Dem Neuropsychol. 2016;10(4):287-295. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 218.Fink KD, Deng P, Torrest A, Stewart H, Pollock K, Gruenloh W, et al. Developing stem cell therapies for juvenile and adult-onset Huntington's disease. Regen Med. 2015;10(5):623-646. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 219.Fink KD, Rossignol J, Crane AT, Davis KK, Bombard MC, Bavar AM, et al. Transplantation of umbilical cord-derived mesenchymal stem cells into the striata of R6/2 mice: Behavioral and neuropathological analysis. Stem Cell Research & Therapy. 2013;4(5):130. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 220.Ebrahimi MJ, Aliaghaei A, Boroujeni ME, Khodagholi F, Meftahi G, Abdollahifar MA, et al. human umbilical cord matrix stem cells reverse oxidative stress-induced cell death and ameliorate motor function and striatal atrophy in rat model of Huntington Disease. Neurotox Res. 2018;34(2):273-284. [DOI] [PubMed] [Google Scholar]
- 221.Vis JC, Verbeek MM, De Waal RM, Ten Donkelaar HJ, Kremer HP. 3-Nitropropionic acid induces a spectrum of Huntington's disease-like neuropathology in rat striatum. Neuropathol Appl Neurobiol. 1999;25(6):513-521. [DOI] [PubMed] [Google Scholar]
- 222.Rossignol J, Boyer C, Lévèque X, Fink KD, Thinard R, Blanchard F, et al. Mesenchymal stem cell transplantation and DMEM administration in a 3NP rat model of Huntington's disease: Morphological and behavioral outcomes. Behav Brain Res. 2011;217(2):369-378. [DOI] [PubMed] [Google Scholar]
- 223.Liu J. Induced pluripotent stem cell-derived neural stem cells: New hope for stroke? Stem Cell Res Ther. 2013;4(5):115. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 224.Mu S, Wang J, Zhou G, Peng W, He Z, Zhao Z, et al. Transplantation of induced pluripotent stem cells improves functional recovery in Huntington's disease rat model. PLoS One. 2014;9(7):e101185. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 225.Sanberg PR, Calderon SF, Giordano M, Tew JM, Norman AB. The quinolinic acid model of Huntington's disease: Locomotor abnormalities. Exp Neurol. 1989;105(1):45-53. [DOI] [PubMed] [Google Scholar]
- 226.Yoon Y, Kim HS, Hong CP, Li E, Jeon I, Park HJ, et al. Neural transplants from human induced pluripotent stem cells rescue the pathology and behavioral defects in a rodent model of Huntington’s disease. Front Neurosci. 2020;14:558204. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 227.Al-Gharaibeh A, Culver R, Stewart AN, Srinageshwar B, Spelde K, Frollo L, et al. Induced pluripotent stem cell-derived neural stem cell transplantations reduced behavioral deficits and ameliorated neuropathological changes in YAC128 mouse model of Huntington's disease. Front Neurosci. 2017;11:628. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 228.Lansita JA, Mease KM, Qiu H, Yednock T, Sankaranarayanan S, Kramer S. Nonclinical development of ANX005: A humanized anti-C1q antibody for treatment of autoimmune and neurodegenerative diseases. Int J Toxicol. 2017;36(6):449-462. [DOI] [PubMed] [Google Scholar]
- 229.Keswani SC. Phase 2a Study of ANX005, a Humanized anti-C1q mAb, in Patients with Huntington disease. Huntington Study Group Annual Conference; October 30; Online2020.
- 230.Cho K. Emerging roles of complement protein C1q in neurodegeneration. Aging Dis. 2019;10(3):652-663. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 231.Annexon I. Single Dose Study of ANX005 in Healthy Volunteers 2017. Available from. https://clinicaltrials.gov/ct2/show/NCT03010046.
- 232.Annexon I. A Clinical Study of ANX005 and IVIG in Subjects With Guillain Barré Syndrome (GBS) 2019. Available from:. https://clinicaltrials.gov/ct2/show/NCT04035135 https://clinicaltrials.gov/ct2/show/NCT04035135.
- 233.Annexon I. An Open Label Study of ANX005 in Subjects With, or at Risk for, Manifest Huntington's Disease 2020 [cited 2021 26 January]. Available from:. https://clinicaltrials.gov/ct2/show/NCT04514367.
- 234.Annexon I. Study of ANX005 in Adults With Amyotrophic Lateral Sclerosis (ALS) 2020. Available from:. https://clinicaltrials.gov/ct2/show/NCT04569435.
- 235.Annexon I. Safety, Tolerability, Pharmacokinetics and Pharmacodynamics of ANX005 in Subjects With Warm Autoimmune Hemolytic Anemia (wAIHA) 2020 [cited 2021 28 January]. Available from. https://clinicaltrials.gov/ct2/show/NCT04691570.
- 236.Islam Z, Papri N, Jahan I, Azad KAK, Kroon H, Humphriss E, et al. Inhibition of C1q, initiator of the classical complement cascade, by ANX005 for the treatment of Guillain-Barré Syndrome. Results from a Phase 1b study. 2020 American Academy of Neurology (AAN) Annual Meeting; 28 April; Toronto, Canada2020.
- 237.Annexon I. Efficacy and Safety of ANX005 in Subjects With Guillain-Barré Syndrome 2021. Available from. https://clinicaltrials.gov/ct2/show/NCT04701164.
- 238.Leonard JE, Fisher TL, Winter LA, Cornelius CA, Reilly C, Smith ES, et al. Nonclinical safety evaluation of VX15/2503, a humanized IgG4 Anti-SEMA4D antibody. Mol Cancer Therapeut. 2015;14(4):964. [DOI] [PubMed] [Google Scholar]
- 239.Smith ES, Jonason A, Reilly C, Veeraraghavan J, Fisher T, Doherty M, et al. SEMA4D compromises blood–brain barrier, activates microglia, and inhibits remyelination in neurodegenerative disease. Neurobiol Dis. 2015;73:254-268. [DOI] [PubMed] [Google Scholar]
- 240.Southwell AL, Franciosi S, Villanueva EB, Xie Y, Winter LA, Veeraraghavan J, et al. Anti-semaphorin 4D immunotherapy ameliorates neuropathology and some cognitive impairment in the YAC128 mouse model of Huntington disease. Neurobiol Dis. 2015;76:46-56. [DOI] [PubMed] [Google Scholar]
- 241.Vaccinex I. A Study in Subjects With Late Prodromal and Early Manifest Huntington's Disease (HD) to Assess the Safety, Tolerability, Pharmacokinetics, and Efficacy of Pepinemab (VX15/2503) (SIGNAL) 2015. Available from. https://clinicaltrials.gov/ct2/show/NCT02481674.
- 242.Rodrigues FB, Wild EJ. Huntington's disease clinical trials corner: August 2018. J Huntingt Dis. 2018;7(3):279-286. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 243.Vaccinex, Inc. Announces Preliminary Data from the SIGNAL Clinical Trial (Investigational Drug VX15/2503 as a Potential Treatment for Huntington’s Disease) [press release]. Rochester, New York 2017.
- 244.Top-Line Results of Phase 2 SIGNAL Study in Huntington’s Disease Support Potential for Cognitive Benefit of Pepinemab [press release]. September 22 2020 2020.
- 245.Vaccinex I. Learnings from the SIGNAL Phase 2 Study of Treatment with Pepinemab Antibody. Huntington Study Group 2020 Medical Conference; 30 October 2020; Virtual2020.
- 246.University of Utah . Anti-SEMA4D Monoclonal Antibody VX15/2503 With Nivolumab or Ipilimumab in Treating Patients With Stage III or IV Melanoma 2018. Available from:. https://www.clinicaltrials.gov/ct2/show/NCT03425461?term=VX15&draw=5&rank=2.
- 247.Vaccinex I. SEMA4D Blockade Safety and Brain Metabolic Activity in Alzheimer's Disease (AD) (SIGNAL-AD) 2020. Available from:. https://www.clinicaltrials.gov/ct2/show/NCT04381468.
- 248.Emory University . VX15/2503 and Immunotherapy in Resectable Pancreatic and Colorectal Cancer 2017. Available from:. https://www.clinicaltrials.gov/ct2/show/NCT03373188 https://www.clinicaltrials.gov/ct2/show/NCT03373188.
- 249.Emory University . VX15/2503 in Combination With Ipilimumab or Nivolumab in Patients With Head and Neck Cancer 2018. Available from:. https://www.clinicaltrials.gov/ct2/show/NCT03690986.
- 250.LaGanke C, Samkoff L, Edwards K, Jung Henson L, Repovic P, Lynch S, et al. Safety/tolerability of the anti-semaphorin 4D Antibody VX15/2503 in a randomized phase 1 trial. Neurol Neuroimmunol Neuroinflamm. 2017;4(4):e367. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 251.Patnaik A, Weiss GJ, Leonard JE, Rasco DW, Sachdev JC, Fisher TL, et al. Safety, pharmacokinetics, and pharmacodynamics of a humanized anti-semaphorin 4D antibody, in a first-in-human study of patients with advanced solid tumors. Clin Cancer Res. 2016;22(4):827-836. [DOI] [PubMed] [Google Scholar]
- 252.Vaccinex I. VX15/2503 in Combination With Avelumab in Advanced Non-small Cell Lung Cancer (CLASSICAL-Lung) 2017. Available from. https://www.clinicaltrials.gov/ct2/show/NCT03268057.
- 253.Khoshnan A, Southwell AL, Bugg CW, Ko JC, Patterson PH. recombinant intrabodies as molecular tools and potential therapeutics for Huntington’s disease. In: Lo DC, Hughes RE, eds Neurobiology of Huntington's Disease: Applications to Drug Discovery. Boca Raton (FL: CRC Press/Taylor & Francis; 2011. [PubMed] [Google Scholar]
- 254.Amaro IA, Henderson LA. An intrabody drug (rAAV6-INT41) reduces the binding of N-terminal Huntingtin fragment(s) to DNA to basal levels in PC12 cells and delays cognitive loss in the R6/2 animal model. J Neurodegn Dis. 2016;2016:7120753. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 255.Zha J, Liu X-M, Zhu J, Liu S-Y, Lu S, Xu P-X, et al. A scFv antibody targeting common oligomeric epitope has potential for treating several amyloidoses. Sci Rep. 2016;6:36631. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 256.Southwell AL, Khoshnan A, Dunn DE, Bugg CW, Lo DC, Patterson PH. Intrabodies binding the proline-rich domains of mutant huntingtin increase its turnover and reduce neurotoxicity. J Neurosci. 2008;28(36):9013-9020. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 257.Miller TW, Zhou C, Gines S, MacDonald ME, Mazarakis ND, Bates GP, et al. A human single-chain Fv intrabody preferentially targets amino-terminal huntingtin fragments in striatal models of Huntington's disease. Neurobiol Dis. 2005;19(1):47-56. [DOI] [PubMed] [Google Scholar]
- 258.Butler DC, Messer A. Bifunctional anti-huntingtin proteasome-directed intrabodies mediate efficient degradation of mutant huntingtin exon 1 protein fragments. PLoS One. 2011;6(12):e29199. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 259.Denis HL, Lauruol F, Cicchetti F. Are immunotherapies for Huntington's disease a realistic option?. Mol Psychiatr. 2019;24:364-377. [DOI] [PubMed] [Google Scholar]
- 260.Positive preclinical in vivo results with AFFiRiS’ antibody mAB C6-17 to treat Huntington’s disease to be presented at the 16th Annual Huntington's Disease Therapeutics Conference [press release]. April 27 2021.
- 261.Bartl S, Oueslati A, Southwell AL, Siddu A, Parth M, David LS, et al. Inhibiting cellular uptake of mutant huntingtin using a monoclonal antibody: Implications for the treatment of Huntington's disease. Neurobiol Dis. 2020;141:104943. [DOI] [PubMed] [Google Scholar]
- 262.Waters S, Tedroff J, Ponten H, Klamer D, Sonesson C, Waters N. Pridopidine: Overview of pharmacology and rationale for its use in Huntington's disease. J Huntingt Dis. 2018;7(1):1-16. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 263.Reilmann R, McGarry A, Grachev ID, Savola JM, Borowsky B, Eyal E, et al. Safety and efficacy of pridopidine in patients with Huntington's disease (PRIDE-HD): A phase 2, randomised, placebo-controlled, multicentre, dose-ranging study. Lancet Neurol. 2019;18(2):165-176. [DOI] [PubMed] [Google Scholar]
- 264.Geva M, Kusko R, Soares H, Fowler KD, Birnberg T, Barash S, et al. Pridopidine activates neuroprotective pathways impaired in Huntington Disease. Hum Mol Genet. 2016;25(18):3975-3987. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 265.Squitieri F, Di Pardo A, Favellato M, Amico E, Maglione V, Frati L. Pridopidine, a dopamine stabilizer, improves motor performance and shows neuroprotective effects in Huntington disease R6/2 mouse model. J Cell Mol Med. 2015;19(11):2540-2548. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 266.Ryskamp D, Wu J, Geva M, Kusko R, Grossman I, Hayden M, et al. The sigma-1 receptor mediates the beneficial effects of pridopidine in a mouse model of Huntington disease. Neurobiol Dis. 2017;97(Pt A):46-59. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 267.Lundin A, Dietrichs E, Haghighi S, Göller M-L, Heiberg A, Loutfi G, et al. Efficacy and safety of the dopaminergic stabilizer Pridopidine (ACR16) in patients with Huntington's disease. Clin Neuropharmacol. 2010;33(5):260-264. [DOI] [PubMed] [Google Scholar]
- 268.de Yebenes JG, Landwehrmeyer B, Squitieri F, Reilmann R, Rosser A, Barker RA, et al. Pridopidine for the treatment of motor function in patients with Huntington's disease (MermaiHD): A phase 3, randomised, double-blind, placebo-controlled trial. Lancet Neurol. 2011;10(12):1049-1057. [DOI] [PubMed] [Google Scholar]
- 269.Huntington Study Group HART Investigators . A randomized, double-blind, placebo-controlled trial of pridopidine in Huntington's disease. Mov Disord. 2013;28(10):1407-1415. [DOI] [PubMed] [Google Scholar]
- 270.Prilenia . A Phase 2, to Evaluating the Safety and Efficacy of Pridopidine Versus Placebo for Symptomatic Treatment in Patients With Huntington's Disease 2013. Available from:. https://clinicaltrials.gov/ct2/show/NCT02006472.
- 271.McGarry A, Leinonen M, Kieburtz K, Geva M, Olanow CW, Hayden M. Effects of Pridopidine on Functional Capacity in Early-Stage Participants from the PRIDE-HD Study. J Huntingt Dis. 2020;9(4):371-380. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 272.Squitieri F, Landwehrmeyer B, Reilmann R, Rosser A, de Yebenes JG, Prang A, et al. One-year safety and tolerability profile of pridopidine in patients with Huntington disease. Neurology. 2013;80(12):1086-1094. [DOI] [PubMed] [Google Scholar]
- 273.McGarry A, Kieburtz K, Abler V, Grachev ID, Gandhi S, Auinger P, et al. Safety and exploratory efficacy at 36 months in open-HART, an open-label extension study of pridopidine in Huntington’s disease. J Huntingt Dis. 2017;6:189-199. [DOI] [PubMed] [Google Scholar]
- 274.McGarry A, Auinger P, Kieburtz K, Geva M, Mehra M, Abler V, et al. Additional safety and exploratory efficacy data at 48 and 60 months from open-HART, an open-label extension study of pridopidine in Huntington disease. J Huntingt Dis. 2020;9:173-184. [DOI] [PubMed] [Google Scholar]
- 275.Prilenia . A Study Evaluating if Pridopidine is Safe, Efficacious, and Tolerable in Patients With Huntington's Disease (Open PRIDE-HD) 2015. Available from:. https://clinicaltrials.gov/ct2/show/NCT02494778.
- 276.Prilenia . PRidopidine's Outcome On Function in Huntington Disease, PROOF- HD 2020. Available from:. https://clinicaltrials.gov/ct2/show/NCT04556656.
- 277.Rodrigues FB, Wild EJ. Clinical trials corner: September 2017. J Huntingt Dis. 2017;6(3):255-263. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 278.Haggiag S, Ruggieri S, Gasperini C. Efficacy and safety of laquinimod in multiple sclerosis: Current status. Therap Adv Neurol Dis. 2013;6(6):343-352. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 279.Brück W, Pförtner R, Pham T, Zhang J, Hayardeny L, Piryatinsky V, et al. Reduced astrocytic NF-κB activation by laquinimod protects from cuprizone-induced demyelination. Acta Neuropathologica. 2012;124(3):411-424. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 280.Filippi M, Rocca MA, Pagani E, De Stefano N, Jeffery D, Kappos L, et al. Placebo-controlled trial of oral laquinimod in multiple sclerosis: MRI evidence of an effect on brain tissue damage. J Neurol Neurosurg Psychiatr. 2014;85(8):851. [DOI] [PubMed] [Google Scholar]
- 281.Garcia-Miralles M, Tan J, Radulescu C, Sidik H, Belinson H, Zach N, et al. Laquinimod treatment improves myelination deficits at the transcriptional and ultrastructural levels in the YAC128 mouse model of Huntington disease. Mol Neurobiol. 2019;56(6):4464-4478. [DOI] [PubMed] [Google Scholar]
- 282.Teva Pharmaceutical Industries . A Clinical Study in Participants With Huntington's Disease (HD) to Assess Efficacy and Safety of Three Oral Doses of Laquinimod (LEGATO-HD) 2014 [updated May 4 2020; cited 2021 17 February]. Available from. https://clinicaltrials.gov/ct2/show/NCT02215616.
- 283.Reilmann R, Gordon MF, Anderson KE, Feigin A, Tabrizi SJ, Leavitt BR, et al. The efficacy and safety results of laquinimod as a treatment for Huntington disease (LEGATO-HD) (S16.007). Neurology. 2019;92(15 suppl ment):S16.007. [Google Scholar]
- 284.Leavitt BR, Reilmann R, Gordon MF, Anderson KE, Feigin A, Tabrizi SJ, et al. eds. Magnetic resonance spectroscopy evaluation of neuronal integrity and astrocytosis in a phase 2 study of laquinimod as a treatment for Huntington disease (LEGATO-HD) International Congress of Parkinson’s Disease and Movement disorders. Nice, France; 2019. [Google Scholar]
- 285.Reilmann R, Gordon MF, Schubert R, Anderson KE, Feigin A, Tabrizi SJ, et al. eds. Quantitative motor (Q-Motor) assessments suggest a beneficial central effect of laquinimod in a phase II study in Huntington disease (LEGATO-HD) International Congress of Parkinson’s Disease and Movement disorders. NiceFrance2019; 2019. [Google Scholar]
- 286.Bachstetter AD, Xing B, de Almeida L, Dimayuga ER, Watterson DM, Van Eldik LJ. Microglial p38α MAPK is a key regulator of proinflammatory cytokine up-regulation induced by toll-like receptor (TLR) ligands or beta-amyloid (Aβ). J Neuroinflammation. 2011;8:79. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 287.Alam JJ. Selective brain-targeted antagonism of p38 MAPKα reduces hippocampal IL-1β levels and improves Morris Water maze performance in aged rats. J Alzheim Dis : JAD. 2015;48(1):219-227. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 288.Haddad JJ. VX-745. Vertex pharmaceuticals. Curr Opin Invest Drugs. 2001;2(8):1070-1076. [PubMed] [Google Scholar]
- 289.EIP Pharma Inc . Within Subject Crossover Study of Cognitive Effects of Neflamapimod in Early-Stage Huntington Disease 2019 [cited 2021 18 February]. Available from. https://clinicaltrials.gov/ct2/show/NCT03980938.
- 290.EIP Pharma Announces Positive Phase 2 Results for Neflamapimod in Mild-to-Moderate Dementia with Lewy bodies (DLB) [press release]. Boston, Massachusetts, 6 October 2020 2020.
- 291.EIP Pharma Announces Clinical Trial Results of REVERSE-SD, a Phase 2b Study of Neflamapimod in Early-stage Alzheimer’s Disease [press release]. Boston, Massachusetts, November 7 2019.
- 292.Pagan F, Hebron M, Valadez EH, Torres-Yaghi Y, Huang X, Mills RR, et al. Nilotinib effects in Parkinson's disease and dementia with lewy bodies. J Parkinsons Dis. 2016;6(3):503-517. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 293.Hebron ML, Lonskaya I, Moussa CEH. Nilotinib reverses loss of dopamine neurons and improves motor behavior via autophagic degradation of α-synuclein in Parkinson's disease models. Hum Mol Genet. 2013;22(16):3315-3328. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 294.Hebron ML, Lonskaya I, Olopade P, Selby ST, Pagan F, Moussa CEH. Tyrosine kinase inhibition regulates early systemic immune changes and modulates the neuroimmune response in α-synucleinopathy. J Clin Cell Immunol. 2014;5:259. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 295.Anderson KE. Nilotinib in Huntington's Disease (Tasigna HD) 2018 [cited 2021 20 February ]. Available from. https://clinicaltrials.gov/ct2/show/NCT03764215.
- 296.Pfizer. Randomized, Placebo Controlled Study Of The Efficacy And Safety Of PF-02545920 In Subjects With Huntington's Disease 2014 [cited 2021 18 February]. Available from: https://clinicaltrials.gov/ct2/show/NCT02197130.
- 297.Beaumont V, Zhong S, Lin H, Xu W, Bradaia A, Steidl E, et al. Phosphodiesterase 10A inhibition improves cortico-basal ganglia function in Huntington's disease models. Neuron. 2016;92(6):1220-1237. [DOI] [PubMed] [Google Scholar]
- 298.Pfizer . Product Pipeline for New Pharmaceutical Drugs 2021 [updated February 2 2021; cited 2021 18 February]. Available from. https://www.pfizer.com/science/drug-product-pipeline.
- 299.Liu GZ, Hou TT, Yuan Y, Hang PZ, Zhao JJ, Sun L, et al. Fenofibrate inhibits atrial metabolic remodelling in atrial fibrillation through PPAR-α/sirtuin 1/PGC-1α pathway. Br J Pharmacol. 2016;173(6):1095-1109. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 300.Johri A, Chandra A, Flint Beal M. PGC-1α, mitochondrial dysfunction, and Huntington's disease. Free Radic Biol Med. 2013;62:37-46. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 301.Brownstein MJ, Simon NG, Long JD, Yankey J, Maibach HT, Cudkowicz M, et al. Safety and tolerability of SRX246, a vasopressin 1a antagonist, in irritable Huntington's disease patients-a randomized phase 2 clinical trial. J Clin Med. 2020;9(11):3682. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 302.McGregor AL, Dysart J, Tingle MD, Russell BR, Kydd RR, Finucane G. Varenicline improves motor and cognitive symptoms in early Huntington's disease. Neuropsychiatric Disease and Treatment. 2016;12:2381-2386. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 303.Sage Therapeutics Announces Planned Progression of SAGE-718 to Phase 2 in Huntington's Disease and Presentations at the 2019 Annual Meeting of the American College of Neuropsychopharmacology (ACNP) [press release]. December 10 2019.
- 304.Sage Therapeutics to Host Sage Science Spotlight: SAGE-718 In Depth [press release]. May 13 2021.
- 305.Huntington Study Group Reach2HD Investigators . Safety, tolerability, and efficacy of PBT2 in Huntington's disease: A phase 2, randomised, double-blind, placebo-controlled trial. Lancet Neurol. 2015;14(1):39-47. [DOI] [PubMed] [Google Scholar]
- 306.Li C, Yuan K, Schluesener H. Impact of minocycline on neurodegenerative diseases in rodents: A meta-analysis. Rev Neurosci. 2013;24(5):553. [DOI] [PubMed] [Google Scholar]
- 307.Bonelli RM, Hödl AK, Hofmann P, Kapfhammer H-P. Neuroprotection in Huntington's disease: A 2-year study on minocycline. Int Clin Psychopharmacol. 2004;19(6):337. [DOI] [PubMed] [Google Scholar]
- 308.Thomas M, Ashizawa T, Jankovic J. Minocycline in Huntington's disease: A pilot study. Mov Disord. 2004;19(6):692-695. [DOI] [PubMed] [Google Scholar]
- 309.Cudkowicz M. A futility study of minocycline in Huntington's disease. Mov Disord. 2010;25(13):2219-2224. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 310.Möller T. Neuroinflammation in Huntington’s disease. J Neural Transm. 2010;117(8):1001-1008. [DOI] [PubMed] [Google Scholar]
- 311.Rocha NP, Ribeiro FM, Furr-Stimming E, Teixeira AL. Neuroimmunology of Huntington’s Disease: Revisiting Evidence from Human Studies. Mediat Inflamm. 2016;2016:8653132. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 312.Stella N. Endocannabinoid signaling in microglial cells. Neuropharmacology. 2009;56:244-253. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 313.Consroe P, Laguna J, Allender J, Snider S, Stern L, Sandyk R, et al. Controlled clinical trial of cannabidiol in Huntington's disease. Pharmacol Biochem Behav. 1991;40(3):701-708. [DOI] [PubMed] [Google Scholar]
- 314.López-Sendón Moreno JL, García Caldentey J, Trigo Cubillo P, Ruiz Romero C, García Ribas G, Alonso Arias M, et al. A double-blind, randomized, cross-over, placebo-controlled, pilot trial with Sativex in Huntington’s disease. J Neurol. 2016;263(7):1390-1400. [DOI] [PubMed] [Google Scholar]
- 315.Masgrau R, Guaza C, Ransohoff RM, Galea E. Should we stop saying ‘glia’ and ‘neuroinflammation’? Trends Mol Med. 2017;23(6):486-500. [DOI] [PubMed] [Google Scholar]
- 316.Matias I, Morgado J, Gomes FCA. Astrocyte heterogeneity: Impact to brain aging and disease. Front Aging Neurosci2019. [DOI] [PMC free article] [PubMed]
- 317.Ransohoff RM. A polarizing question: Do M1 and M2 microglia exist? Nat Neurosci. 2016;19(8):987-991. [DOI] [PubMed] [Google Scholar]
- 318.Sturrock A, Leavitt BR. The clinical and genetic features of Huntington disease. J Geriatr Psychiatr Neurol. 2010;23(4):243-259. [DOI] [PubMed] [Google Scholar]
- 319.Frank S. Treatment of Huntington's disease. Neurotherapeutics. 2014;11(1):153-160. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 320.André VM, Cepeda C, Levine MS. Dopamine and glutamate in Huntington's disease: A balancing act. CNS Neuroscience & Therapeutics. 2010;16(3):163-178. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 321.Chen J, Wang E, Cepeda C, Levine M. Dopamine imbalance in Huntington's disease: A mechanism for the lack of behavioral flexibility. Front Neurosci. 2013;7. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 322.Miller BR, Bezprozvanny I. Corticostriatal circuit dysfunction in Huntington’s disease: Intersection of glutamate, dopamine and calcium. Future Neurol. 2010;5(5):735-756. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 323.Zuccato C, Cattaneo E. Role of brain-derived neurotrophic factor in Huntington's disease. Prog Neurobiol. 2007;81(5):294-330. [DOI] [PubMed] [Google Scholar]
- 324.Xie Y, Hayden MR, Xu B. BDNF overexpression in the forebrain rescues Huntington's disease phenotypes in YAC128 mice. J Neurosci. 2010;30(44):14708-14718. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 325.Erta M, Quintana A, Hidalgo J. Interleukin-6, a major cytokine in the central nervous system. Int J Biol Sci. 2012;8(9):1254-1266. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 326.Costa V, Scorrano L. Shaping the role of mitochondria in the pathogenesis of Huntington's disease. EMBO J. 2012;31(8):1853-1864. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 327.Chaudhary RK, Patel KA, Patel MK, Joshi RH, Roy I. Inhibition of Aggregation of Mutant Huntingtin by Nucleic Acid Aptamers In Vitro and in a Yeast Model of Huntington's Disease. Mol Ther. 2015;23(12):1912-1926. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 328.Zhang X, Smith DL, Meriin AB, Engemann S, Russel DE, Roark M, et al. A potent small molecule inhibits polyglutamine aggregation in Huntington's disease neurons and suppresses neurodegeneration in vivo. Proc Natl Acad Sci USA. 2005;102(3):892. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 329.Senior K. Synaptic activity could hold the key to therapy for Huntington disease. Nat Rev Neurol. 2010;6(1):1.20238450 [Google Scholar]
- 330.Sepers MD, Raymond LA. Mechanisms of synaptic dysfunction and excitotoxicity in Huntington's disease. Drug Discov Today. 2014;19(7):990-996. [DOI] [PubMed] [Google Scholar]
- 331.Marques Sousa C, Humbert S. Huntingtin: Here, there, everywhere! J Huntingt Dis. 2013;2(4):395-403. [DOI] [PubMed] [Google Scholar]
- 332.Carroll JB, Bates GP, Steffan J, Saft C, Tabrizi SJ. Treating the whole body in Huntington's disease. Lancet Neurol. 2015;14(11):1135-1142. [DOI] [PubMed] [Google Scholar]
- 333.van der Burg JMM, Björkqvist M, Brundin P. Beyond the brain: Widespread pathology in Huntington's disease. Lancet Neurol. 2009;8(8):765-774. [DOI] [PubMed] [Google Scholar]
- 334.Kong G, Ellul S, Narayana VK, Kanojia K, Ha HTT, Li S, et al. An integrated metagenomics and metabolomics approach implicates the microbiota-gut-brain axis in the pathogenesis of Huntington's disease. Neurobiol Dis. 2021;148:105199. [DOI] [PubMed] [Google Scholar]
- 335.Du G, Dong W, Yang Q, Yu X, Ma J, Gu W, et al. Altered gut microbiota related to inflammatory responses in patients with Huntington's disease. Front Immunol. 2020;11:603594. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 336.Stan TL, Soylu-Kucharz R, Burleigh S, Prykhodko O, Cao L, Franke N, et al. Increased intestinal permeability and gut dysbiosis in the R6/2 mouse model of Huntington’s disease. Sci Rep. 2020;10(1):18270. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 337.Wasser CI, Mercieca E-C, Kong G, Hannan AJ, McKeown SJ, Glikmann-Johnston Y, et al. Gut dysbiosis in Huntington's disease: Associations among gut microbiota, cognitive performance and clinical outcomes. Brain Commun. 2020;2(2):fcaa110. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 338.Kong G, Cao KL, Judd LM, Li S, Renoir T, Hannan AJ. Microbiome profiling reveals gut dysbiosis in a transgenic mouse model of Huntington's disease. Neurobiol Dis. 2020;135:104268. [DOI] [PubMed] [Google Scholar]
- 339.McColgan P, Tabrizi SJ. Huntington's disease: A clinical review. Eur J Neurol. 2018;25(1):24-34. [DOI] [PubMed] [Google Scholar]
- 340.Harper SQ. Progress and challenges in RNA interference therapy for Huntington disease. Arch Neurol. 2009;66(8):933-938. [DOI] [PubMed] [Google Scholar]
- 341.Meghan S, Neil A. Frontiers in neuroscience RNA- and DNA- based therapies for Huntington’s disease. In: Lo DC, Hughes RE, eds Neurobiology of Huntington's Disease: Applications to Drug Discovery. Boca Raton (FL: CRC Press/Taylor & Francis; 2011:. [Google Scholar]





