Next Article in Journal
Loxl2 and Loxl3 Paralogues Play Redundant Roles during Mouse Development
Next Article in Special Issue
Mutations, Genes, and Phenotypes Related to Movement Disorders and Ataxias
Previous Article in Journal
Comparative Analyses of Medicinal Chemistry and Cheminformatics Filters with Accessible Implementation in Konstanz Information Miner (KNIME)
Previous Article in Special Issue
Valosin Containing Protein (VCP): A Multistep Regulator of Autophagy
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Recent Insight into the Genetic Basis, Clinical Features, and Diagnostic Methods for Neuronal Ceroid Lipofuscinosis

1
Students’ Scientific Society, Department of Pediatric Neurology, Faculty of Medical Sciences in Katowice, Medical University of Silesia, 40-752 Katowice, Poland
2
Pediatric Neurology Department, Faculty of Medical Sciences in Katowice, Medical University of Silesia, 40-752 Katowice, Poland
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2022, 23(10), 5729; https://0-doi-org.brum.beds.ac.uk/10.3390/ijms23105729
Submission received: 14 April 2022 / Revised: 15 May 2022 / Accepted: 17 May 2022 / Published: 20 May 2022
(This article belongs to the Special Issue Molecular Genetics in Neurodegenerative Disorders)

Abstract

:
Neuronal ceroid lipofuscinoses (NCLs) are a group of rare, inherited, neurodegenerative lysosomal storage disorders that affect children and adults. They are traditionally grouped together, based on shared clinical symptoms and pathological ground. To date, 13 autosomal recessive gene variants, as well as one autosomal dominant gene variant, of NCL have been described. These genes encode a variety of proteins, whose functions have not been fully defined; most are lysosomal enzymes, transmembrane proteins of the lysosome, or other organelles. Common symptoms of NCLs include the progressive loss of vision, mental and motor deterioration, epileptic seizures, premature death, and, in rare adult-onset cases, dementia. Depending on the mutation, these symptoms can vary, with respect to the severity and onset of symptoms by age. Currently, all forms of NCL are fatal, and no curative treatments are available. Herein, we provide an overview to summarize the current knowledge regarding the pathophysiology, genetics, and clinical manifestation of these conditions, as well as the approach to diagnosis.

1. Introduction

The term neuronal ceroid lipofuscinoses (NCL) refers to a group of autosomal recessive neurodegenerative disorders, presenting with myoclonic epilepsy, psychomotor delay, progressive loss of vision, and early death [1,2,3,4]. It is worth noting that NCL are the most common cause of childhood dementia [3,5,6]. Cardinal to NCL pathology is the toxic levels of protein aggregates in the central nervous system (CNS) [4,5,7], which, more specifically, are aggregates of lipopigments (lipofuscin) within lysosomes [4,8,9]. They are the most prevalent neurodegenerative disorders of childhood, with an incidence in the USA estimated at 1.6–2.4/100,000 [2,4,10,11,12], while, in Scandinavian countries, the incidence varies between: 2–2.5/100,000 in Denmark [2,10,11], 2.2/100,000 in Sweden [2,10,11], 3.9/100,000 in Norway [2,10,11], 4.8/100,000 in Finland [2,10,12], and 7/100,000 in Iceland [2,10]. The rarity of incidence and scarcity of the disease models have limited comprehensive understanding of the pathological factors that lead to disease progression. To date, 13 autosomal recessive, as well as one autosomal dominant gene, variants of NCL have been described, each involving functional defects in lysosomal protein [3,4,13,14,15,16]. Aggregated lysosomal lipofuscin affects the neuronal cytoskeleton and cellular trafficking, resulting in neuronal loss and pathological glial proliferation and activation [4,7]. More specifically, defective proteins (corresponding genes designated CLN for “ceroid lipofuscinosis, neuronal”) include secreted lysosomal proteins (CLN1/PPT1, CLN2/TPP1, CLN5, CLN10/CTSD, CLN13/CTSF, and CLN11/GRN) or membrane proteins (CLN3, CLN7/MFSD8, and CLN12/ATP13A2) [4,17,18,19,20]. Two of them are the membrane proteins of the endoplasmic reticulum (CLN6 and CLN8). Other NCL proteins are cytoplasmic (CLN4/DNAJC5 and CLN14/KCTD7), which peripherally associate with cell membranes [4,17,18,19,20].
NCLs are grouped on pathological grounds, due to the common presence of the neuronal and extra-neuronal accumulations of autofluorescent pigments, despite the different underlying biochemical etiologies [2,3,21]. In all forms of NCL, lipopigment storage material accumulates in macrophages, neurons, and some somatic tissues, including vascular endothelial cells and smooth muscle cells [3,22]. Neuronal loss is profound and extensive in most patients with NCL, leading to cortical gray matter atrophy, cerebellar atrophy, and secondary ventricular enlargement, all of which progress throughout the course of the disease [3,22,23]. The degree of atrophy and ventricular enlargement varies with the form of NCL [3,22,23]. The prominent activation of microglia and astrocytes precedes (and perhaps causes) neuronal loss, accurately predicting its distribution [3,24,25,26].
NCL is classified into five primary types, as shown in Table 1.
The most common NCL types are the CLN1 (also known as “Infantile NCL”), CLN2 (“Late Infantile”), CLN3 (“Juvenile NCL”), and CLN6 (“variant LINCL”) diseases [4,17,18,19]. Table 2 shows all known types of neuronal ceroid lipofuscinoses, together with the genes whose mutations are responsible for their occurrence. Table 2 also contains the age of onset of the disease and the proteins whose functions are impaired.
In 1998, a database was established that includes published mutations and sequence changes in genes causing NCL, together with unpublished data, included with permission. All mutation details can be found in the publicly available NCL mutation database at the following Internet address: www.ucl.ac.uk/ncl-disease (accessed on 16 May 2022).

2. Genetics and Pathophysiology

2.1. Animal Models in NCL Pathology

The development of various animal models has provided scientists with a range of tools for studying the effects of the mutations responsible for NCL. In NCL, as in many other neurodegenerative diseases, there is usually a subset of tissues or cell populations that are selectively susceptible to pathogenic agents [5,33,34,35].
Animal models (especially the widely used mouse models of NCL) and their characteristics allow researchers to develop disease phenotypes at the organismal level, as they largely reproduce certain disease-relevant phenotypes, with obvious species limitations [5,34,35,36,37,38,39,40,41,42,43,44]. With the advent of more accurate and precise technologies to generate mouse models with specific mutations, attempts have been made to replicate the most common human disease-causing mutations in mice, which would allow for a better representation and understanding of the pathological changes observed in human NCL [5,45,46,47,48]. Naturally, in addition to mouse models, larger animal models of NCL have been developed in dogs [5,49,50,51,52] and sheep [5,53]. These models have proved very useful for understanding the anatomical and pathophysiological spread of disease, as they more closely mimic the human disease and provide a means to test the delivery and dosing of therapeutic agents in a way that is not possible in mice [5].
The use of CRISPR-Cas9 has enabled the generation of a detailed and genetically comprehensive sheep model of CLN1 disease [5,54], as well as a porcine model of CLN3 disease [5,55]. In all likelihood, we can expect similar models to be developed for other forms of NCL [5].

2.2. Anatomical Regions Affected by NCL Pathology

A constant characteristic sign in NCL is cerebral and cerebellar atrophy co-occurring with enlargement of the lateral ventricles in the brain [5]. It should be noted, however, that atrophy is not a uniform process; thus, some regions are affected much earlier than others [5,23,56,57,58,59]. Studies in mouse models have shown that the thalamus and cerebellum are particularly vulnerable regions in various forms of NCL, and somatosensory regions of the cortex are affected earlier and more severely than motor regions [5,42,60,61,62,63]. Indicative changes include marked glial activation, accumulation of autofluorescent storage material (AFSM), and neuronal loss, as well as marked loss of interneurons in the cortex and hippocampus. The stage and progression of pathology has also been determined in larger animal models of NCL, confirming that these regions may, indeed, be clinically relevant [5,51,52,64,65,66]. Patients with NCL are often characterised by sensory and motor deficits [5,13]. Spinal cord involvement, as demonstrated in CLN1, may be responsible for this [67]. A study in animal models, as well as autopsy studies in humans, suggest that spinal cord pathology occurs with high frequency in many types of NCL [5,22,51,54,55,68].
Most patients with NCLs suffer from progressive loss of vision [3,5,13]. This has necessitated a detailed examination of the visual pathway, which has shown marked retinal degeneration in these patients [13,69,70]. In many forms of NCL, based on animal models, it is also often accompanied by optic nerve degeneration [5,71,72,73,74,75,76,77]. Furthermore, mouse models have been shown to exhibit significant pathology in central visual pathways within the dorsolateral geniculate nucleus of the thalamus and visual cortex in many forms of NCL [42,60,61,62], and it is likely that other retinal receptive nuclei are also affected. Another issue with mouse models is that rodent species rely much more on sensory information from the eyeballs than on visual information [78], which may explain why the somatosensory pathways through the posterior ventral thalamic nucleus to the barrel-field cortex are so severely impaired in the mouse models of NCL [61,62,63,79,80], even more so than the central visual pathways [42,60,61,62].
In NCLs, as in many other known storage diseases, findings provide evidence of significant cardiac pathology from both clinical observations and experimental studies [5,81,82,83]. These are best understood in CLN3 disease and also include evidence of autonomic nervous system dysfunction [84]. Furthermore, research reports also refer to the intestinal problems occurring in children with NCL [5,13].
Since most of the proteins in the NCL are widely expressed in different tissues and cell types, in all likelihood, we can suspect that the deficiency of these proteins also affects other organ systems in the body (including outside the CNS) [19,85,86]. Therefore, treating only the regions in the brain occupied by the disease process may not be sufficient to achieve therapeutic success. Making the identification of other, unexpected sites where pathology develops seem important. While the involvement of the brain by the disease process appears to be the primary concern, other disease effects, associated with other locations of disease development (although not to the same extent as the brain), will potentially also require therapeutic intervention. Treating these previously overlooked pathologies may provide additional clinical benefits, improving quality of life [5].

2.3. Cell Types in NCL Pathology

NCL affects different cell types [87,88,89,90]. Dysfunction and loss of the number of properly functioning neurons are characteristic features, bearing in mind that neuronal population types differ in their sensitivity to being affected [5,88]. One type of neuron that is more susceptible to damage in NCL is the interneuron population. The greatest loss of this population occurs in the thalamus, cerebral cortex, and hippocampus, where they account for the majority of neurons lost in the early stages of the disease [60,61,79].
Their greater vulnerability to damage may be due to their properties, mainly electrophysiological and bioenergetic [91]. What is clear in NCLs is that AFSMs are deposited in lysosomes, and their function is impaired; as it seems, the preserved normal lysosomal function is critical for the proper functioning of interneurons [47,92]. Interestingly, interneuron populations have also been shown to play an important role in the neurodegenerative pathways of various diseases, such as epilepsy [93], Alzheimer’s [94], amyotrophic lateral sclerosis (ALS) [95], and Parkinson’s [96].
Despite the fact that AFSMs accumulate in many cell types, NCLs have always been considered neuronal diseases, which is also reflected in their name, i.e., “neuronal ceroid lipofuscinoses” [5,83]. Nowadays, an early and significant effect of NCLs on other cells, which are also part of the CNS, is increasingly pointed out. These include astrocytes and microglia, which are observed in a state of activation, both biochemically and histologically [5,48,60,61,97]. Microglia activation generally occurs at a stage immediately preceding neuronal loss, and its location has a greater predictive value, in relation to the site of neuronal loss, than the site of storage material appearance [3,98]. Certainly, the question of whether glial activation contributes to neuronal population loss (which is a suggested pathophysiological pathway in other lysosomal or neurodegenerative disorders) still needs to be clarified [5,99,100,101]. Primary cell culture experiments investigating the properties of glial cells derived from mouse models of CLN1 and CLN3 have shown that astrocytes and microglia have significant functional and morphological defects, although they vary widely between these forms of NCL [5,25,26]. In co-culture systems, astrocytes and microglia have been shown to damage or negatively affect the survival of both healthy and mutant neurons [5,25]. These experiences highlight the role that glial cells appear to play in the pathogenesis of NCL. In addition to the innate immune changes evidenced by glial activation in the CNS, there is also evidence for a general humoral immune response in NCL [102,103], with a possible autoimmune component, particularly in CLN3 [5,104]. Authors should discuss the results and how they can be interpreted, from the perspective of previous studies and the working hypotheses. The findings and their implications should be discussed in the broadest context possible. Future research directions may also be highlighted.

2.4. Cellular Model

Mutations in the NCL involve the endo-lysosomal system. Lysosomes are responsible for recycling endogenous and internalized molecules, which are involved in maintaining tissue homeostasis [105,106]. They are also known to be involved in many cellular processes, such as nutrient sensing, calcium and bio-metal homeostasis, cell growth axonal transport, and synaptic homeostasis [105,106]. Because of this, autophagy function, which is essential for normal cellular homeostasis, is impaired in the NCL. Despite the disclosure of these disorders in various forms of NCL, the exact mechanism that leads to the consequent changes that cause cell death still remains unknown [47,80,107,108,109]. Similar disturbances in the autophagy mechanism have been revealed in other lysosomal disorders [110,111].
Animal models have also shown changes in synaptic vesicle density, exocytosis, and electrophysiological changes, which are indicative of synaptic dysfunction [62,65,108,112,113,114]. Despite the evidence that lysosomes are transported into the synapse to facilitate synaptic remodelling, pre-synaptic autophagy, synaptic vesicle sorting, and overall synaptic homeostasis [5,115,116,117], the exact mechanisms causing NCL to affect the lysosomal system, leading to synaptic dysfunction, are still not understood [5,115,116,117]. However, the question of synaptic dysfunction remains essential for a comprehensive understanding of the entire pathomechanism, mainly due to the fact that it often precedes the stage of neuronal loss. It is highly likely that this is due to altered conductivity within the neurons, as well as the presence of other signalling defects within cells [5,118]. Disease involvement of the (already mentioned) astrocytes and microglia is also one of the postulated mechanisms leading to loss of synaptic function [5,118,119,120,121].
The discovery of the cellular mechanisms responsible for controlling the expression of many endo-lysosomal proteins is undoubtedly one of the major achievements of molecular biology. These include the discovery of the coordinated lysosomal expression and regulation (CLEAR) transcription initiation site, the EB transcription factor (TFEB), and its phosphorylation by mTORC1. Of particular relevance to the pathophysiology of NCLs was the demonstration of dysregulation of this pathway in many NCL types [5,122,123,124,125,126]. Preclinical studies in a mouse model provided evidence that the aforementioned dysregulation may be a viable therapeutic target [127,128], mainly because an upregulation of TFEB increases the overall cellular clearance, which may bypass the existing lysosomal dysfunction [127,128].

2.5. Genotype–Phenotype Correlations

All of the groups of neuronal ceroid lipofuscinoses are monogenic disorders, so each represents a distinct disease entity. Genes whose mutations are responsible for the occurrence of NCL encode seemingly unrelated proteins, including the soluble lysosomal enzymes and membrane proteins located in various organelles, including the lysosome [3,19,129,130]. All types of NCL are inherited autosomal recessively, except neuronal ceroid lipofuscinosis type 4 (CLN4) [3,129]. For most NCLs, there is a recognizable classical disease phenotype associated with the complete loss of gene function, due to intracellular mislocalization or degradation of the mutant protein [129]. In addition, there are forms of the disease that may have a more chronic course, and some of the predicted classical phenotypes may not be present, as a result of milder mutations that do not cause complete arrest of protein function [3,24]. There are examples of mutations associated with a specific phenotype, such as a missense mutation in CLN8 or a 1 kb intragenic deletion that underlies the most common form of NCL, i.e., juvenile CLN3 disease [3,24,129]. The most common mutations are a 1 kb deletion in CLN3 and two mutations in CLN2 [3,23,24].

3. Clinical Features

The clinical picture in which NCL was diagnosed in the past was the result of careful analysis of a compilation of case reports; nowadays, natural history reports, regarding the course of the disease in patients, continue to be used to advance knowledge of NCL and to prepare clinical trials [28,131,132,133,134,135]. In the clinical setting, the age and onset of symptoms continue to play a key role in suspecting NCL and, ultimately, pursuing confirmatory genetic testing for the disease. All patients with NCL, except those with the rare congenital form (neuronal ceroid lipofuscinosis type 10 (CLN10)), have normal psychomotor development before the onset of first symptoms [3,13]. In most patients, the time of onset of the first noticeable symptoms of the disease is during childhood; however, in some, it can be as long as 60 years or more. NCL is a collection of disparate subtypes, but typical clinical manifestations of NCL include progressive loss of vision, dementia, psychomotor stunting, inability to reach normal developmental milestones and/or developmental regression, behavioral problems, progressive brain atrophy, seizures, cognitive decline, and motor skills [3,16,19,24,130,136,137,138].
Cognitive decline in young children occurs rapidly, while adolescents and adults show a much slower rate, allowing for normal life and education. Cognitive decline occurs as a two-phase problem [139]. Early in the disease, children’s developmental trajectory slows; they acquire new learning more slowly than their peers but often remain within the normal range for some time [139]. They go on to plateau and then begin to lose the cognitive competences they have acquired during the earliest months and/or years of their lives [139]. This is often the time at which the diagnostic investigations are triggered. In CLN2 disease, however, many children show a delayed acquisition of cognitive skills at a very early stage; their profiles are characteristically uneven, with delays being more evident in the domain of expressive language, compared with motor skills [139]. In CLN3, the rate of disease progression is commonly slow, and children can attend school, with support for the intellectual, visual, and, sometimes, behavioral difficulties. There is no evidence, to date, of an uneven developmental profile early in life and before the onset of progressive vision loss [139].
Symptoms of the disease can also affect other regions, often not directly related to the central nervous system. This is the case with CLN3, where cardiac involvement is often described in adolescent and adult patients [84,140]. This problem is amenable to treatment by implanting a pacemaker, when the patient’s psychomotor performance has improved [3,139].
Among the first symptoms in the classic (age of onset 6–24 months) and late (age of onset 2–5 years) infantile phenotypes is a slowing of psychomotor development, quickly followed by developmental arrest, gradual loss of acquired psychomotor skills, onset of epilepsy, and, finally, loss of vision [13]. Unfortunately, these disorders, particularly the regression of psychomotor abilities, are often misdiagnosed as side effects of the drugs used (for example, antiepileptic drugs), which delays the diagnosis of NCL and, by further exacerbating developmental defects, worsens the child’s condition. On the other hand, in the juvenile phenotype (age of onset 5–7 years), the first symptoms are usually the progressive loss of vision, followed by the development of dementia and behavioral changes, and, finally, the loss of motor skills and epilepsy in early adolescence. In the recessive adult phenotype (over 16 years of age), also referred to as Kufs disease, visual loss is usually absent, and patients develop progressive myoclonic epilepsy (type A) or dementia with motor impairment (type B), which usually begins between 16 and 50 years of age, but most commonly around age 30 [3,24].
The delayed development of expressive language, also known as expressive language disorder, is revealed as the first symptom indicating the deterioration of psychomotor function in 83% of patients with classic late-onset CLN2 disease—this may be an indication to screen the child for CLN2 and allow for early diagnosis [3,134,141].
One of the persistent symptoms of NCL also includes epilepsy. This is very troublesome, as it is characterized by treatment resistance in almost all patients. The use of more than two antiepileptic drugs is ineffective, as it only exacerbates their side effects, without controlling or reducing the number of seizures [3,13]. In the pharmacotherapy of epilepsy occurring in the CLN2 and CLN3 subtypes, lamotrigine and valproic acid are the recommended drugs [3,142,143], whereas drugs such as phenytoin, gabapentin, vigabatrin, or carbamazepine may cause an increase in myoclonic seizures in these patients [3,142,143]. The frequency and severity of epileptic seizures is particularly high in late infantile CLN2, which continues until the late stages of the disease [3,141]. The situation is different, for example, in infantile neuronal ceroid lipofuscinosis type 1 (CLN1), where seizure frequency tends to decrease in the later stages of the disease, or in patients with classic juvenile CLN3, where seizures are rare and only moderate severity occurs in the later stages of the disease [3,141,142,143]. As the disease progresses, the benefits and harms of specific medications should be reconsidered, as these drugs, while tolerable and effective in the earlier stages of the disease, may prove ineffective after progression and cause only side effects [143].
Typical neurological manifestations of NCL also include motor symptoms, such as ataxia, dysphagia, myoclonus, chorea, tremors, and dystonia [3]. These especially occur in the classic and late infantile phenotypes [3,13]. In addition, parkinsonism is also present, especially in juvenile CLN3 disease, and some stereotypic movements have also been described in various types of NCL with late infantile and juvenile onset [3,13]. Table 3 shows the clinical characteristics of the different subtypes of neuronal ceroid lipofuscinoses.

4. Diagnosis

Many known childhood neurodegenerative diseases have similar symptoms, so the delayed diagnosis of NCL is unfortunately common [16,144]. Definitive diagnosis can be problematic in infants and young children, in whom a thorough neurological and ophthalmic evaluation requires a skilled and knowledgeable specialist. NCL occurs after a period of apparently normal development, despite the absence of a functional protein that is important for brain function; thus, it must be concluded that there is a small therapeutic window in which effective interventions can halt and/or prevent disease progression. Consequently, early diagnosis is crucial for optimal therapeutic outcomes, and prospective studies must also include those patients with only modest disease progression. Consequently, knowledge of the clinical features, which unfortunately are highly nonspecific, seems to be a necessity for making a suspected diagnosis and carrying out the diagnostic process, in order to be able to definitively confirm the diagnosis as soon as possible and to incorporate the treatment adapted to it.
However, there is no single, universally valid diagnostic algorithm for neuronal ceroid lipofuscinoses. The gold standard diagnostic tool for NCL is genetic testing, in order to detect and confirm the presence of the mutated gene; however, these tests are expensive and may take months to analyze and provide results. It is important to keep in mind, however, that this is not one of the first tests of choice, neuronal ceroid lipofuscinosis is a very rare disorder; before the clinician can guide his diagnostic process toward NCL, it is necessary to first rule out other, much more common potential causes of the condition that the patient is experiencing. Tests that are used to diagnose NCL in a patient with suspected clinical symptoms include enzyme tests, neuroimaging studies (e.g., magnetic resonance imaging or computed tomography), visual evoked potentials, electroencephalography (EEG), and other electrophysiological studies [16,145,146,147]. Genome analysis will certainly identify the known pathogenic variants; however, in some cases, detection of the common disease-causing mutations is not confirmed. In this case, we expand the diagnosis with biochemical assays. When biochemical impairment is found, the diagnostic search for changes in cryptic gene regions (introns, untranslated regions, etc.) is extended [139].

4.1. Genetic Testing

Technological advances and new DNA testing technologies, mainly concerning genomic sequencing, now make it possible to study multiple genes in a single step, regardless of presentation [3]. Whole exome, whole genome, or direct Sanger sequencing are the most reliable analytical tools for the accurate identification of mutations in patients [16,129,147,148] and should be performed in such rare diseases because overlap with other neurodegenerative disorders is blurred. These methods are particularly helpful for patients with clinical variants and/or unusual or novel mutations. The significant heterogeneity of this group of patients, due to the vast differences in age of onset, clinical presentation, severity, and disease course, makes genetic testing the only universal tool at our disposal, with nearly 100% sensitivity, that clinicians and researchers can use to identify the specific genetic mutations that may cause protein dysfunction and confirm the diagnosis of neuronal ceroid lipofuscinoses [16,129,147,148].
Currently, in the diagnostic process, genes whose mutations are responsible for the occurrence of NCL are part of the panels designed to examine the genes underlying a larger group of syndromically and nonsyndromically inherited epilepsies. Some common mutations can be tested using DNA-based assays. This can expedite the earlier diagnosis of NCL before the onset of other symptoms, as well as provide a genetic diagnosis for milder clinical phenotypes or variants. Because DNA sequencing leads to the description of many genetic variants, the genetic cause of atypical disease will become clearer in some cases. Some patients who were previously diagnosed with NCL may turn out to have atypical forms of other diseases and vice versa. Detection of the carrier is not possible by histology, and it is unreliable by enzymatic analysis; mutation analysis should always be relied upon.
There is a mutation database that catalogs all known variant mutations of the NCL genes, which includes nearly 446 disease-causing mutations, and it should be noted that new disease-causing gene mutations are still being discovered [16,20,129].

4.2. Enzyme and Microscopic Assays

Biochemical assays for the activity of four lysosomal enzymes that are mutated in various forms of the disease (PPT1, TPP1, CTSF, and cathepsin D (encoded by CTSD)) can rapidly confirm their deficiencies and are among the conclusive methods for diagnosing CLN1, CLN2, and CLN10, respectively [16,20,139,149,150,151,152,153]. These tests can be performed using saliva obtained from the patient, a blood sample, or a dried blood spot [3,20]. Enzyme tests should always be used in cases with an unusual appearance or later onset, and all diagnoses should be supported by DNA sequencing and mutation analysis, if possible [20,130]. For PPT1 testing, patient samples, including fibroblasts, leukocytes, amniocytes, dried blood spots, and chorionic villi, are mixed with a synthetic substrate that is hydrolyzed by the active PPT1 enzyme to release a detectable fluorophore [16,20,152,153]. The absence of a fluorophore confirms CLN1 disease. Similarly, TPP1 and CTSD activity assays use fluorogenic synthetic substrates conjugated to Ala-Ala-Phe and conjugated to hemoglobin, respectively, in which the active enzyme cleaves the substrate to activate the fluorophore [16,20,154,155,156]. Each of these tests can be considered rapid and reliable tools for assisting clinicians in the initial diagnosis of the disorder, as well as in distinguishing its potential subtypes [16,20]. In adults with nonspecific mental, motor, or behavioral disorders, in whom NCL is suspected, the first line of testing includes the just-mentioned enzymatic tests for PPT1, TTP1, CTSD, and CTSF, which, when normal, should prompt the clinician to perform ultrastructural testing [2,16,20]. If storage material is present, genetic testing should be initiated for autosomal recessive (CLN6, CLN11/GRN, and CLN13/CTSF) and dominant (CLN4/DNAJC5) NCL; if negative, all other NCL genes should be tested [2,16,20]. The occurrence of disease in a newborn with severe epilepsy and microcephaly should suggest CLN10 disease as a possible diagnosis. Enzymatic testing for CTSD (CLN10) should be the first step. If negative, further or concurrent enzyme testing for PPT1 and TTP1 should be performed before more invasive biochemical testing.
In young children (>6 months) with unexplained epilepsy and developmental arrest, CLN1 and CLN2 are most likely to be affected. If enzyme tests for PPT1 and TTP1 are negative and electron microscopy shows typical storage material, genetic testing for CLN5, CLN6, CLN7/MFSD8, CLN8, and CLN14/KCTD7 should be considered [2].
In countries where genetic and biochemical testing are not readily available, skin biopsy samples can also be obtained, with the lipopigment accumulation assessed, which is a pathological feature of these disorders [16,23,157]. Microscopic analysis of ultrastructural patterns of cellular deposits helps divide patients into the possible subtypes of NCL. Lipopigment morphotypes usually correlate strongly with the genotype [16,35,158]; however, the genotype does not always coincide with the clinical presentation of the disease. For this reason, storage deposits can be used as a confirmation of neuronal ceroid lipofuscinosis subtypes, rather than as a diagnostic tool, with different morphotypes indicating different forms of the disease [2,16,23]. Ultrastructural examination of a skin biopsy or blood sample can be helpful in confirming the NCL disease subtype for atypical forms that are not enzyme deficient or genetically undiagnosed. Extracerebral storage is readily detectable in childhood NCL, but not necessarily in NCL occurring in adulthood [21,24].
In addition, in patients suspected of having classic juvenile CLN3 disease, such as a school-aged child with rapid visual loss between 4 and 7 years of age, one should look for vacuolated lymphocytes. This feature, characteristic of CLN3, can be assessed by a simple blood smear, and it can be helpful in the differential diagnosis of patients [2,16,19,138]. In contrast, if no lymphocyte vacuolization is found and the PPT1, TPP1, and CTSD tests are negative, it is advisable to perform the aforementioned skin biopsy to assess whether typical storage material for NCL is present. If so, genetic testing for CLN5, CLN6, CLN7/MFSD8, CLN8, and CLN12/ATP13A2 is indicated [2].

4.3. Neuroimaging

Brain MRI findings may appear normal in the early stages of disease or show some nonspecific signs, such as periventricular intensity changes in early CLN2 disease [3,159]. Although MRI is not sensitive or specific for early diagnosis, it is an excellent tool for objectively monitoring the progression of brain changes, especially with changes in resolution and processing, and neuroimaging techniques, such as diffusion tensor imaging, allow for the assessment of white matter pathway disorganization and atrophy [3,23,54,55,56,57,58,59]. Magnetic resonance imaging in advanced stages of the disease can mainly visualize cortical gray matter atrophy, cerebellar atrophy, and secondary ventricular enlargement [3,23,54,55,56,57,58,59]. The results in patients with CLN2 disease have shown that the loss of cortical gray matter volume with age can be a sensitive biomarker for monitoring disease progression [2,58,159].

4.4. EEG Monitoring and Other Electrophisiological Studies

EEG can be helpful in the early diagnosis of some NCLs. For example, characteristic posterior spike waves after low-frequency photostimulation have been reported in more than 90% of patients with early-stage CLN2 disease [3,160]. The finding of posterior spike waves after photostimulation in a child under 5 years of age with new onset seizures should be a reason to initiate testing for CLN2 disease [3,160,161]. Photosensitivity with low-frequency stimulation has also been described in patients with CLN6 (neuronal ceroid lipofuscinosis type 6) disease, especially in adults with CLN6 (type A) disease. [3,137,162,163]. In forms of NCL that progress rapidly, such as infantile CLN1 disease, early abnormalities disappear as neurons die, leading to a characteristic flat EEG in advanced disease [3,13]. Importantly, in NCL, due to the underlying neurometabolic disease, taking antiepileptic drugs does not lead to seizure control and the normalization of EEG results [3,13], possibly because the cells have died (or are dying), which itself leads to seizures. EEG monitoring may be useful in detecting signs of encephalitis, which may be a rare consequence of antiepileptic drugs, such as valproate, in advanced, late-infantile CLN2 disease [3,143] or a theoretical allergic reaction to new treatments [3].
Another of the electrophysiological studies that are crucial in rare diseases, such as NCL, is photoparoxysmal response (PPR). It was first described in the 1980s [164,165], and subsequent studies have confirmed this neurophysiological finding [165,166,167,168,169]. The PPR seen in the NCLs is characterized by an occipital spike-and-waves response to the photic stimuli [165]. However, it should also be highlighted that patients do not have the same susceptibility to PPR in all NCL variants—it has been reported more frequently in CLN2, CLN5, and CLN6 diseases [165]. In CLN2 disease, PPR has been reported from 27 to 93% of patients [165,166,168,169,170]. In a series of 14 CLN2 cases, serial EEG showed PPR in 93% of patients [165,166]. PPR was evident from the first EEG in 43% of patients; it was documented at low (1–3 Hz) stimulation frequencies in 69% of patients and took the form of a flash-per-flash response in 69% of patients [165,166]. Interestingly, in advanced stages of the disease, PPR was associated with massive myoclonic jerks [165,166]. In approximately half of patients (54%), PPR disappears over time [164,165]. Changes in PPR over time reflect changes in the grey matter caused by the progression of neurodegenerative disease [165,166]. It is also worth mentioning that PPR is one of the earliest pieces of evidence of a neurodegenerative disorder, even before the onset of cognitive and motor regression [165,171].
To date, little is known about changes in visual evoked potentials (VEPs) during the course of disease in NCLs [165]. It is known that so-called giant VEPs appear as the disease progresses. They are abnormally broad and of high amplitude; their presence is a marker of cortical hyperactivity, similar to PPR [165,167]. Abnormal VEPs were found in approximately 75% of CLN2 patients [165,170]. The majority of patients (89%) had an early reversal of the high-amplitude pattern in the VEP, while a small number of patients showed a bifid waveform, which is associated with central scotoma, indicative of maculopathy or macular pathway dysfunction, such as optic nerve atrophy [165,172]. As the disease progresses further, the amplitude of the VEP decreases, as it does in other neurodegenerative disorders [165].
In addition to visual evoked potentials, there are also somatosensory evoked potentials (SEP), which are less well-studied in NCL than VEPs. There are sporadic cases across different forms of NCLs [149,165,167,173,174], where it is highlighted that the presence of high-amplitude evoked potentially giant SEPs, which are the expressions of cortical hyperexcitability, due to neuronal degeneration [165,167].
Progressive loss of vision is one of the classical symptoms of NCL. It affects all forms; in the large majority of cases, it occurs as one of the early clinical signs [139,165]. Retinal structures, visual pathways, and visual cortices are affected [139]. Both ganglionic neurons and receptor cells (cones and rods) are involved [139]. For this reason, an electroretinogram (ERG) is also one of the electrophysiological studies [139,165]. It is a tool that allows for the monitoring of retinal involvement, and its use gradually disappears as the disease progresses, depending on the variants of NCL [165,175]. The use of ERG might be implemented in the future to evaluate the efficacy of experimental treatments with intravitreal therapies [165].

4.5. Prenatal Diagnosis

Prenatal diagnosis may be offered to families with prior NCL disease. Preimplantation genetic diagnosis [21,174] or a combination of enzyme assay and mutation analysis, perhaps with ultrastructural examination of chorionic samples obtained at 12–15 weeks gestation, may provide a rapid diagnosis [3,20].
However, genetic testing and enzyme activity assays are now the standard. These methods are readily available in most developed countries and can reliably be performed prenatally using amniotic fluid or fetal cells [16,23,149,175,176,177].
In Table 4, we have presented the subtypes of neuronal ceroid lipofuscinosis and their possible deviations in the results of tests that are performed during the diagnostic process (neuroimaging, EEG monitoring, and visual and microscopic examinations).

5. Conclusions

NCLs include a group of rare, fatal, neurodegenerative, lysosomal storage disorders. They predominantly affect the pediatric population, but there are also cases with disease onset in adulthood. Specific mutations cause the onset of the disease and often correlate with the clinical phenotype. Some variability in the clinical phenotype can, therefore, be explained by different levels of residual protein function. However, variability between families, and even between siblings, shows that co-inheritance of other genetic variations may affect the disease phenotype. The typical clinical manifestations include progressive loss of vision, mental and motor deterioration, epileptic seizures, behavioural problems, cognitive decline, and dementia. All of these significantly reduce the patients’ quality of life. Through functional impairment and the resulting dependence on caregivers and facilities, the exclusion of patients from society continues to progress. Understanding the molecular basis of NCL seems to play a key role, as knowing the exact cause and pathomechanisms can lead to effective causal treatment. This is extremely important, as current treatment options are only symptomatic and focus on delaying disease progression. It is worth noting the international cooperation in creating a database of patients and mutations, which is of great importance for accessing information on each type of NCL, as well as for planning future clinical trials.

Author Contributions

Conceptualization was performed by J.P. Methodology was performed by K.K., S.K. and J.P. Software by K.K., S.K. and J.P. Validation was performed by K.K., S.K. and J.P. Formal analysis was performed by J.P. Investigation was performed by K.K., S.K. and J.P. Resources were provided by K.K., S.K. and J.P. Data curation was performed by K.K., S.K. and J.P. Writing—original draft preparation was performed by K.K., S.K. and J.P. Writing—review and editing was performed by J.P. Visualization was performed by K.K., S.K. and J.P. Supervision was performed by J.P. Project administration was performed by J.P. Funding acquisition was performed by J.P. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Chabrol, B.; Caillaud, C.; Minassian, B. Neuronal ceroid lipofuscinoses. Handb. Clin. Neurol. 2013, 113, 1701–1706. [Google Scholar] [CrossRef] [PubMed]
  2. Nita, D.A.; Mole, S.E.; Minassian, B.A. Neuronal ceroid lipofuscinoses. Epileptic Disord. 2016, 18, 73–88. [Google Scholar] [CrossRef]
  3. Mole, S.E.; Anderson, G.; Band, H.A.; Berkovic, S.F.; Cooper, J.D.; Kleine Holthaus, S.M.; McKay, T.R.; Medina, D.L.; Rahim, A.A.; Schulz, A.; et al. Clinical challenges and future therapeutic approaches for neuronal ceroid lipofuscinosis. Lancet Neurol. 2019, 18, 107–116. [Google Scholar] [CrossRef]
  4. Specchio, N.; Ferretti, A.; Trivisano, M.; Pietrafusa, N.; Pepi, C.; Calabrese, C.; Livadiotti, S.; Simonetti, A.; Rossi, P.; Curatolo, P.; et al. Neuronal Ceroid Lipofuscinosis: Potential for Targeted Therapy. Drugs 2021, 81, 101–123. [Google Scholar] [CrossRef] [PubMed]
  5. Nelvagal, H.R.; Lange, J.; Takahashi, K.; Tarczyluk-Wells, M.A.; Cooper, J.D. Pathomechanisms in the neuronal ceroid lipofuscinoses. Biochim. Biophys. Acta-Mol. Basis Dis. 2020, 1866, 165570. [Google Scholar] [CrossRef]
  6. Haltia, M.; Goebel, H.H. The neuronal ceroid-lipofuscinoses: A historical introduction. Biochim. Biophys. Acta-Mol. Basis Dis. 2013, 1832, 1795–1800. [Google Scholar] [CrossRef] [Green Version]
  7. Platt, F.M. Emptying the stores: Lysosomal diseases and therapeutic strategies. Nat. Rev. Drug Discov. 2018, 17, 133–150. [Google Scholar] [CrossRef]
  8. Jalanko, A.; Braulke, T. Neuronal ceroid lipofuscinoses. Biochim. Biophys. Acta-Mol. Cell Res. 2009, 1793, 697–709. [Google Scholar] [CrossRef] [Green Version]
  9. Glees, P.; Hasan, M. Lipofuscin in neuronal aging and diseases. Norm. Pathol. Anat. 1976, 32, 1–68. [Google Scholar]
  10. Uvebrant, P.; Hagberg, B. Neuronal ceroid lipofuscinoses in Scandinavia: Epidemiology and clinical pictures. Neuropediatrics 1997, 28, 6–8. [Google Scholar] [CrossRef]
  11. Mole, S.; Haltia, M. The neuronal ceroid-lipofuscinoses (Batten disease). In Rosenberg’s Molecular and Genetic Basis of Neurological and Psychiatric Disease, 5th ed.; Rosenberg, R., Pascual, J., Eds.; Academic Press: Boston, MA, USA, 2015; pp. 793–808. [Google Scholar]
  12. Cardona, F.; Rosati, E. Neuronal ceroid-lipofuscinoses in Italy: An epidemiological study. Am. J. Med. Genet. 1995, 57, 142–143. [Google Scholar] [CrossRef] [PubMed]
  13. Schulz, A.; Kohlschütter, A.; Mink, J.; Simonati, A.; Williams, R. NCL diseases—Clinical perspectives. Biochim. Biophys. Acta-Mol. Basis Dis. 2013, 1832, 1801–1806. [Google Scholar] [CrossRef] [Green Version]
  14. Donsante, A.; Boulis, N.M. Progress in gene and cell therapies for the neuronal ceroid lipofuscinoses. Expert Opin. Biol. Ther. 2018, 18, 755–764. [Google Scholar] [CrossRef] [PubMed]
  15. Rakheja, D.; Bennett, M.J. Neuronal ceroid-lipofuscinoses. In Metabolic Diseases: Foundations of Clinical Management, Genetics, and Pathology; Gilbert-Barness, E., Barness, L.A., Farrell, P.M., Eds.; IOS Press BV: Amsterdam, The Netherlands, 2017; pp. 499–510. [Google Scholar]
  16. Johnson, T.B.; Cain, J.T.; White, K.A.; Ramirez-Montealegre, D.; Pearce, D.A.; Weimer, J.M. Therapeutic landscape for Batten disease: Current treatments and future prospects. Nat. Rev. Neurol. 2019, 15, 161–178. [Google Scholar] [CrossRef]
  17. Mole, S.E.; Williams, R.E.; Goebel, H.H. Correlations between genotype, ultrastructural morphology and clinical phenotype in the neuronal ceroid lipofuscinoses. Neurogenetics 2005, 6, 107–126. [Google Scholar] [CrossRef] [PubMed]
  18. Getty, A.L.; Pearce, D.A. Interactions of the proteins of neuronal ceroid lipofuscinosis: Clues to function. Cell. Mol. Life Sci. 2011, 68, 453–474. [Google Scholar] [CrossRef] [Green Version]
  19. Cárcel-Trullols, J.; Kovács, A.D.; Pearce, D.A. Cell biology of the NCL proteins: What they do and don’t do. Biochim. Biophys. Acta-Mol. Basis Dis. 2015, 1852 Pt B, 2242–2255. [Google Scholar] [CrossRef] [Green Version]
  20. Warrier, V.; Vieira, M.; Mole, S.E. Genetic basis and phenotypic correlations of the neuronal ceroid lipofusinoses. Biochim Biophys Acta-Mol. Basis Dis. 2013, 1832, 1827–1830. [Google Scholar] [CrossRef] [Green Version]
  21. Morsy, A.; Carmona, A.V.; Trippier, P.C. Patient-Derived Induced Pluripotent Stem Cell Models for Phenotypic Screening in the Neuronal Ceroid Lipofuscinoses. Molecules 2021, 26, 6235. [Google Scholar] [CrossRef]
  22. Anderson, G.W.; Goebel, H.H.; Simonati, A. Human pathology in NCL. Biochim Biophys Acta-Mol. Basis Dis. 2013, 1832, 1807–1826. [Google Scholar] [CrossRef] [Green Version]
  23. Radke, J.; Stenzel, W.; Goebel, H.H. Human NCL Neuropathology. Biochim Biophys Acta-Mol. Basis Dis. 2015, 1852 Pt B, 2262–2266. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  24. Berkovic, S.F.; Staropoli, J.F.; Carpenter, S.; Oliver, K.L.; Kmoch, S.; Anderson, G.W.; Damiano, J.A.; Hildebrand, M.S.; Sims, K.B.; Cotman, S.L.; et al. Diagnosis and misdiagnosis of adult neuronal ceroid lipofuscinosis (Kufs disease). Neurology 2016, 87, 579–584. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  25. Parviainen, L.; Dihanich, S.; Anderson, G.W.; Wong, A.M.; Brooks, H.R.; Abeti, R.; Rezaie, P.; Lalli, G.; Pope, S.; Heales, S.J.; et al. Glial cells are functionally impaired in juvenile neuronal ceroid lipofuscinosis and detrimental to neurons. Acta Neuropathol. Commun. 2017, 5, 74. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  26. Lange, J.; Haslett, L.J.; Lloyd-Evans, E.; Pocock, J.M.; Sands, M.S.; Williams, B.P.; Cooper, J.D. Compromised astrocyte function and survival negatively impact neurons in infantile neuronal ceroid lipofuscinosis. Acta Neuropathol. Commun. 2018, 6, 74. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  27. Siintola, E.; Partanen, S.; Stromme, P.; Haapanen, A.; Haltia, M.; Maehlen, J.; Lehesjoki, A.E.; Tyynela, J. Cathepsin D deficiency underlies congenital human neuronal ceroid-lipofuscinosis. Brain 2006, 129, 1438–1445. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  28. Mink, J.W.; Augustine, E.F.; Adams, H.R.; Marshall, F.J.; Kwon, M.J. Classification and natural history of the neuronal ceroid lipofuscinoses. J. Child Neurol. 2013, 28, 1101–1105. [Google Scholar] [CrossRef] [Green Version]
  29. Marques, A.R.A.; di Spiezio, A.; Thiessen, N.; Schmidt, L.; Grotzinger, J.; Lullmann-Rauch, R.; Damme, M.; Storck, S.E.; Pietrzik, C.U.; Fogh, J.; et al. Enzyme replacement therapy with recombinant pro-CTSD (cathepsin D) corrects defective proteolysis and autophagy in neuronal ceroid lipofuscinosis. Autophagy 2020, 16, 811–825. [Google Scholar] [CrossRef]
  30. Arrant, A.E.; Onyilo, V.C.; Unger, D.E.; Roberson, E.D. Progranulin gene therapy improves lysosomal dysfunction and microglial pathology associated with frontotemporal dementia and neuronal ceroid lipofuscinosis. J. Neurosci. 2018, 38, 2341–2358. [Google Scholar] [CrossRef] [Green Version]
  31. Williams, R.E.; Mole, S.E. New nomenclature and classification scheme for the neuronal ceroid lipofuscinoses. Neurology 2012, 79, 183–191. [Google Scholar] [CrossRef]
  32. Mukherjee, A.B.; Appu, A.P.; Sadhukhan, T.; Casey, S.; Mondal, A.; Zhang, Z.; Bagh, M.B. Emerging new roles of the lysosome and neuronal ceroid lipofuscinoses. Mol. Neurodegener. 2019, 14, 4. [Google Scholar] [CrossRef] [Green Version]
  33. Platt, F.M.; d’Azzo, A.; Davidson, B.L.; Neufeld, E.F.; Tifft, C.J. Lysosomal storage diseases. Nat. Rev. Dis. Primers 2019, 5, 34. [Google Scholar] [CrossRef] [Green Version]
  34. Nixon, R.A.; Yang, D.S.; Lee, J.H. Neurodegenerative lysosomal disorders: A continuum from development to late age. Autophagy 2008, 4, 590–599. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  35. Fu, H.; Hardy, J.; Duff, K.E. Selective vulnerability in neurodegenerative diseases. Nat. Neurosci. 2018, 21, 1350–1358. [Google Scholar] [CrossRef] [PubMed]
  36. Mitchison, H.M.; Bernard, D.J.; Greene, N.D.; Cooper, J.D.; Junaid, M.A.; Pullarkat, R.K.; de Vos, N.; Breuning, M.H.; Owens, J.W.; Mobley, W.C.; et al. Targeted disruption of the Cln3 gene provides a mouse model for Batten disease. The Batten Mouse Model Consortium [corrected]. Neurobiol. Dis. 1999, 6, 321–334. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  37. Koike, M.; Nakanishi, H.; Saftig, P.; Ezaki, J.; Isahara, K.; Ohsawa, Y.; Schulz-Schaeffer, W.; Watanabe, T.; Waguri, S.; Kametaka, S.; et al. Cathepsin D deficiency induces lysosomal storage with ceroid lipofuscin in mouse CNS neurons. J. Neurosci. 2000, 20, 6898–6906. [Google Scholar] [CrossRef] [PubMed]
  38. Schultheis, P.J.; Fleming, S.M.; Clippinger, A.K.; Lewis, J.; Tsunemi, T.; Giasson, B.; Dickson, D.W.; Mazzulli, J.R.; Bardgett, M.E.; Haik, K.L.; et al. Atp13a2-deficient mice exhibit neuronal ceroid lipofuscinosis, limited α-synuclein accumulation and age-dependent sensorimotor deficits. Hum. Mol. Genet. 2013, 22, 2067–2082. [Google Scholar] [CrossRef] [Green Version]
  39. Gupta, P.; Soyombo, A.A.; Atashband, A.; Wisniewski, K.E.; Shelton, J.M.; Richardson, J.A.; Hammer, R.E.; Hofmann, S.L. Disruption of PPT1 or PPT2 causes neuronal ceroid lipofuscinosis in knockout mice. Proc. Natl. Acad. Sci. USA 2001, 98, 13566–13571. [Google Scholar] [CrossRef] [Green Version]
  40. Gao, H.; Boustany, R.M.; Espinola, J.A.; Cotman, S.L.; Srinidhi, L.; Antonellis, K.A.; Gillis, T.; Qin, X.; Liu, S.; Donahue, L.R.; et al. Mutations in a novel CLN6-encoded transmembrane protein cause variant neuronal ceroid lipofuscinosis in man and mouse. Am. J. Hum. Genet. 2002, 70, 324–335. [Google Scholar] [CrossRef] [Green Version]
  41. Kopra, O.; Vesa, J.; von Schantz, C.; Manninen, T.; Minye, H.; Fabritius, A.L.; Rapola, J.; van Diggelen, O.P.; Saarela, J.; Jalanko, A.; et al. A mouse model for Finnish variant late infantile neuronal ceroid lipofuscinosis, CLN5, reveals neuropathology associated with early aging. Hum. Mol. Genet. 2004, 13, 2893–2906. [Google Scholar] [CrossRef] [Green Version]
  42. Sleat, D.E.; Wiseman, J.A.; El-Banna, M.; Kim, K.H.; Mao, Q.; Price, S.; Macauley, S.L.; Sidman, R.L.; Shen, M.M.; Zhao, Q.; et al. A mouse model of classical late-infantile neuronal ceroid lipofuscinosis based on targeted disruption of the CLN2 gene results in a loss of tripeptidyl-peptidase I activity and progressive neurodegeneration. J. Neurosci. 2004, 24, 9117–9126. [Google Scholar] [CrossRef] [Green Version]
  43. Ahmed, Z.; Sheng, H.; Xu, Y.F.; Lin, W.L.; Innes, A.E.; Gass, J.; Yu, X.; Wuertzer, C.A.; Hou, H.; Chiba, S.; et al. Accelerated lipofuscinosis and ubiquitination in granulin knockout mice suggest a role for progranulin in successful aging. Am. J. Pathol. 2010, 177, 311–324. [Google Scholar] [CrossRef] [PubMed]
  44. Tang, C.H.; Lee, J.W.; Galvez, M.G.; Robillard, L.; Mole, S.E.; Chapman, H.A. Murine cathepsin F deficiency causes neuronal lipofuscinosis and late-onset neurological disease. Mol. Cell. Biol. 2006, 26, 2309–2316. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  45. Bouchelion, A.; Zhang, Z.; Li, Y.; Qian, H.; Mukherjee, A.B. Mice homozygous for c.451C>T mutation in Cln1 gene recapitulate INCL phenotype. Ann. Clin. Transl. Neurol. 2014, 1, 1006–1023. [Google Scholar] [CrossRef] [PubMed]
  46. Miller, J.N.; Kovács, A.D.; Pearce, D.A. The novel Cln1R151X mouse model of infantile neuronal ceroid lipofuscinosis (INCL) for testing nonsense suppression therapy. Hum. Mol. Genet. 2015, 24, 185–196. [Google Scholar] [CrossRef] [PubMed]
  47. Brandenstein, L.; Schweizer, M.; Sedlacik, J.; Fiehler, J.; Storch, S. Lysosomal dysfunction and impaired autophagy in a novel mouse model deficient for the lysosomal membrane protein Cln7. Hum. Mol. Genet. 2016, 25, 777–791. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  48. Geraets, R.D.; Langin, L.M.; Cain, J.T.; Parker, C.M.; Beraldi, R.; Kovacs, A.D.; Weimer, J.M.; Pearce, D.A. A tailored mouse model of CLN2 disease: A nonsense mutant for testing personalized therapies. PLoS ONE 2017, 12, e0176526. [Google Scholar] [CrossRef]
  49. Sanders, D.N.; Farias, F.H.; Johnson, G.S.; Chiang, V.; Cook, J.R.; O’Brien, D.P.; Hofmann, S.L.; Lu, J.Y.; Katz, M.L. A mutation in canine PPT1 causes early onset neuronal ceroid lipofuscinosis in a Dachshund. Mol. Genet. Metab. 2010, 100, 349–356. [Google Scholar] [CrossRef] [Green Version]
  50. Katz, M.L.; Farias, F.H.; Sanders, D.N.; Zeng, R.; Khan, S.; Johnson, G.S.; O’Brien, D.P. A missense mutation in canine CLN6 in an Australian shepherd with neuronal ceroid lipofuscinosis. J. Biomed. Biotechnol. 2011, 2011, 198042. [Google Scholar] [CrossRef] [Green Version]
  51. Hirz, M.; Drögemüller, M.; Schänzer, A.; Jagannathan, V.; Dietschi, E.; Goebel, H.H.; Hecht, W.; Laubner, S.; Schmidt, M.J.; Steffen, F.; et al. Neuronal ceroid lipofuscinosis (NCL) is caused by the entire deletion of CLN8 in the Alpenländische Dachsbracke dog. Mol. Genet. Metab. 2017, 120, 269–277. [Google Scholar] [CrossRef]
  52. Katz, M.L.; Rustad, E.; Robinson, G.O.; Whiting, R.; Student, J.T.; Coates, J.R.; Narfstrom, K. Canine neuronal ceroid lipofuscinoses: Promising models for preclinical testing of therapeutic interventions. Neurobiol. Dis. 2017, 108, 277–287. [Google Scholar] [CrossRef]
  53. Palmer, D.N.; Neverman, N.J.; Chen, J.Z.; Chang, C.T.; Houweling, P.J.; Barry, L.A.; Tammen, I.; Hughes, S.M.; Mitchell, N.L. Recent studies of ovine neuronal ceroid lipofuscinoses from BARN, the Batten Animal Research Network. Biochim. Biophys. Acta-Mol. Basis Dis. 2015, 1852 Pt B, 2279–2286. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  54. Eaton, S.L.; Proudfoot, C.; Lillico, S.G.; Skehel, P.; Kline, R.A.; Hamer, K.; Rzechorzek, N.M.; Clutton, E.; Gregson, R.; King, T.; et al. CRISPR/Cas9 mediated generation of an ovine model for infantile neuronal ceroid lipofuscinosis (CLN1 disease). Sci. Rep. 2019, 9, 9891. [Google Scholar] [CrossRef] [PubMed]
  55. Johnson, T.B.; Sturdevant, D.A.; White, K.A.; Drack, A.V.; Bhattarai, S.; Rogers, C.; Cooper, J.D.; Pearce, D.A.; Weimer, J.M. Characterization of a novel porcine model of CLN3-batten disease. Mol. Genet. Metab. 2019, 126, S81. [Google Scholar] [CrossRef]
  56. Oswald, M.J.; Palmer, D.N.; Kay, G.W.; Barwell, K.J.; Cooper, J.D. Location and connectivity determine GABAergic interneuron survival in the brains of South Hampshire sheep with CLN6 neuronal ceroid lipofuscinosis. Neurobiol. Dis. 2008, 32, 50–65. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  57. Autti, T.; Raininko, R.; Santavuori, P.; Vanhanen, S.L.; Poutanen, V.P.; Haltia, M. MRI of neuronal ceroid lipofuscinosis. II. Postmortem MRI and histopathological study of the brain in 16 cases of neuronal ceroid lipofuscinosis of juvenile or late infantile type. Neuroradiology 1997, 39, 371–377. [Google Scholar] [CrossRef]
  58. Löbel, U.; Sedlacik, J.; Nickel, M.; Lezius, S.; Fiehler, J.; Nestrasil, I.; Kohlschütter, A.; Schulz, A. Volumetric Description of Brain Atrophy in Neuronal Ceroid Lipofuscinosis 2: Supratentorial Gray Matter Shows Uniform Disease Progression. Am. J. Neuroradiol. 2016, 37, 1938–1943. [Google Scholar] [CrossRef] [Green Version]
  59. Baker, E.H.; Levin, S.W.; Zhang, Z.; Mukherjee, A.B. MRI Brain Volume Measurements in Infantile Neuronal Ceroid Lipofuscinosis. Am. J. Neuroradiol. 2017, 38, 376–382. [Google Scholar] [CrossRef]
  60. Pontikis, C.C.; Cotman, S.L.; MacDonald, M.E.; Cooper, J.D. Thalamocortical neuron loss and localized astrocytosis in the Cln3Deltaex7/8 knock-in mouse model of Batten disease. Neurobiol. Dis. 2005, 20, 823–836. [Google Scholar] [CrossRef]
  61. Kielar, C.; Maddox, L.; Bible, E.; Pontikis, C.C.; Macauley, S.L.; Griffey, M.A.; Wong, M.; Sands, M.S.; Cooper, J.D. Successive neuron loss in the thalamus and cortex in a mouse model of infantile neuronal ceroid lipofuscinosis. Neurobiol. Dis. 2007, 25, 150–162. [Google Scholar] [CrossRef] [Green Version]
  62. Von Schantz, C.; Kielar, C.; Hansen, S.N.; Pontikis, C.C.; Alexander, N.A.; Kopra, O.; Jalanko, A.; Cooper, J.D. Progressive thalamocortical neuron loss in Cln5 deficient mice: Distinct effects in Finnish variant late infantile NCL. Neurobiol. Dis. 2009, 34, 308–319. [Google Scholar] [CrossRef] [Green Version]
  63. Kuronen, M.; Lehesjoki, A.E.; Jalanko, A.; Cooper, J.D.; Kopra, O. Selective spatiotemporal patterns of glial activation and neuron loss in the sensory thalamocortical pathways of neuronal ceroid lipofuscinosis 8 mice. Neurobiol. Dis. 2012, 47, 444–457. [Google Scholar] [CrossRef] [PubMed]
  64. Perentos, N.; Martins, A.Q.; Cumming, R.J.; Mitchell, N.L.; Palmer, D.N.; Sawiak, S.J.; Morton, A.J. An EEG Investigation of Sleep Homeostasis in Healthy and CLN5 Batten Disease Affected Sheep. J. Neurosci. 2016, 36, 8238–8249. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  65. Amorim, I.S.; Mitchell, N.L.; Palmer, D.N.; Sawiak, S.J.; Mason, R.; Wishart, T.M.; Gillingwater, T.H. Molecular neuropathology of the synapse in sheep with CLN5 Batten disease. Brain Behav. 2015, 5, e00401. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  66. Russell, K.N.; Mitchell, N.L.; Anderson, N.G.; Bunt, C.R.; Wellby, M.P.; Melzer, T.R.; Barrell, G.K.; Palmer, D.N. Computed tomography provides enhanced techniques for longitudinal monitoring of progressive intracranial volume loss associated with regional neurodegeneration in ovine neuronal ceroid lipofuscinoses. Brain Behav. 2018, 8, e01096. [Google Scholar] [CrossRef]
  67. Shyng, C.; Nelvagal, H.R.; Dearborn, J.T.; Tyynelä, J.; Schmidt, R.E.; Sands, M.S.; Cooper, J.D. Synergistic effects of treating the spinal cord and brain in CLN1 disease. Proc. Natl. Acad. Sci. USA 2017, 114, E5920–E5929. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  68. Katz, M.L.; Tecedor, L.; Chen, Y.; Williamson, B.G.; Lysenko, E.; Wininger, F.A.; Young, W.M.; Johnson, G.C.; Whiting, R.E.; Coates, J.R.; et al. AAV gene transfer delays disease onset in a TPP1-deficient canine model of the late infantile form of Batten disease. Sci. Transl. Med. 2015, 7, 313ra180. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  69. Bennett, M.J.; Rakheja, D. The neuronal ceroid-lipofuscinoses. Dev. Disabil. Res. Rev. 2013, 17, 254–259. [Google Scholar] [CrossRef]
  70. Dulz, S.; Wagenfeld, L.; Nickel, M.; Richard, G.; Schwartz, R.; Bartsch, U.; Kohlschütter, A.; Schulz, A. Novel morphological macular findings in juvenile CLN3 disease. Br. J. Ophthalmol. 2016, 100, 824–828. [Google Scholar] [CrossRef]
  71. Leinonen, H.; Keksa-Goldsteine, V.; Ragauskas, S.; Kohlmann, P.; Singh, Y.; Savchenko, E.; Puranen, J.; Malm, T.; Kalesnykas, G.; Koistinaho, J.; et al. Retinal Degeneration in a Mouse Model of CLN5 Disease Is Associated with Compromised Autophagy. Sci. Rep. 2017, 7, 1597. [Google Scholar] [CrossRef] [Green Version]
  72. Preising, M.N.; Abura, M.; Jäger, M.; Wassill, K.H.; Lorenz, B. Ocular morphology and function in juvenile neuronal ceroid lipofuscinosis (CLN3) in the first decade of life. Ophthalmic Genet. 2017, 38, 252–259. [Google Scholar] [CrossRef]
  73. Volz, C.; Mirza, M.; Langmann, T.; Jägle, H. Further Characterization of the Predominant Inner Retinal Degeneration of Aging Cln3Δex7/8 Knock-In Mice. Adv. Exp. Med. Biol. 2018, 1074, 403–411. [Google Scholar] [CrossRef] [PubMed]
  74. Groh, J.; Stadler, D.; Buttmann, M.; Martini, R. Non-invasive assessment of retinal alterations in mouse models of infantile and juvenile neuronal ceroid lipofuscinosis by spectral domain optical coherence tomography. Acta Neuropathol. Commun. 2014, 2, 54. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  75. Follo, C.; Ozzano, M.; Mugoni, V.; Castino, R.; Santoro, M.; Isidoro, C. Knock-down of cathepsin D affects the retinal pigment epithelium, impairs swim-bladder ontogenesis and causes premature death in zebrafish. PLoS ONE 2011, 6, e21908. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  76. Mirza, M.; Volz, C.; Karlstetter, M.; Langiu, M.; Somogyi, A.; Ruonala, M.O.; Tamm, E.R.; Jägle, H.; Langmann, T. Progressive retinal degeneration and glial activation in the CLN6 (nclf) mouse model of neuronal ceroid lipofuscinosis: A beneficial effect of DHA and curcumin supplementation. PLoS ONE 2013, 8, e75963. [Google Scholar] [CrossRef]
  77. Jankowiak, W.; Brandenstein, L.; Dulz, S.; Hagel, C.; Storch, S.; Bartsch, U. Retinal Degeneration in Mice Deficient in the Lysosomal Membrane Protein CLN7. Investig. Ophthalmol. Vis. Sci. 2016, 57, 4989–4998. [Google Scholar] [CrossRef] [Green Version]
  78. Gavornik, J.P.; Bear, M.F. Higher brain functions served by the lowly rodent primary visual cortex. Learn. Mem. 2014, 21, 527–533. [Google Scholar] [CrossRef] [Green Version]
  79. Morgan, J.P.; Magee, H.; Wong, A.; Nelson, T.; Koch, B.; Cooper, J.D.; Weimer, J.M. A murine model of variant late infantile ceroid lipofuscinosis recapitulates behavioral and pathological phenotypes of human disease. PLoS ONE 2013, 8, e78694. [Google Scholar] [CrossRef] [Green Version]
  80. Thelen, M.; Damme, M.; Schweizer, M.; Hagel, C.; Wong, A.M.; Cooper, J.D.; Braulke, T.; Galliciotti, G. Disruption of the autophagy-lysosome pathway is involved in neuropathology of the nclf mouse model of neuronal ceroid lipofuscinosis. PLoS ONE 2012, 7, e35493. [Google Scholar] [CrossRef]
  81. Ostergaard, J.R.; Rasmussen, T.B.; Mølgaard, H. Cardiac involvement in juvenile neuronal ceroid lipofuscinosis (Batten disease). Neurology 2011, 76, 1245–1251. [Google Scholar] [CrossRef]
  82. Ashwini, A.; D’Angelo, A.; Yamato, O.; Giordano, C.; Cagnotti, G.; Harcourt-Brown, T.; Mhlanga-Mutangadura, T.; Guo, J.; Johnson, G.S.; Katz, M.L. Neuronal ceroid lipofuscinosis associated with an MFSD8 mutation in Chihuahuas. Mol. Genet. Metab. 2016, 118, 326–332. [Google Scholar] [CrossRef]
  83. Katz, M.L.; Johnson, G.C.; Leach, S.B.; Williamson, B.G.; Coates, J.R.; Whiting, R.; Vansteenkiste, D.P.; Whitney, M.S. Extraneuronal pathology in a canine model of CLN2 neuronal ceroid lipofuscinosis after intracerebroventricular gene therapy that delays neurological disease progression. Gene Ther. 2017, 24, 215–223. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  84. Ostergaard, J.R. Paroxysmal sympathetic hyperactivity in Juvenile neuronal ceroid lipofuscinosis (Batten disease). Auton. Neurosci. 2018, 214, 15–18. [Google Scholar] [CrossRef] [PubMed]
  85. Fabritius, A.L.; Vesa, J.; Minye, H.M.; Nakano, I.; Kornblum, H.; Peltonen, L. Neuronal ceroid lipofuscinosis genes, CLN2, CLN3 and CLN5 are spatially and temporally co-expressed in a developing mouse brain. Exp. Mol. Pathol. 2014, 97, 484–491. [Google Scholar] [CrossRef] [PubMed]
  86. Minye, H.M.; Fabritius, A.L.; Vesa, J.; Peltonen, L. Data on characterizing the gene expression patterns of neuronal ceroid lipofuscinosis genes: CLN1, CLN2, CLN3, CLN5 and their association to interneuron and neurotransmission markers: Parvalbumin and Somatostatin. Data Brief. 2016, 8, 741–749. [Google Scholar] [CrossRef] [PubMed]
  87. Shacka, J.J. Mouse models of neuronal ceroid lipofuscinoses: Useful pre-clinical tools to delineate disease pathophysiology and validate therapeutics. Brain Res. Bull. 2012, 88, 43–57. [Google Scholar] [CrossRef]
  88. Cooper, J.D.; Tarczyluk, M.A.; Nelvagal, H.R. Towards a new understanding of NCL pathogenesis. Biochim. Biophys. Acta-Mol. Basis Dis. 2015, 1852 Pt B, 2256–2261. [Google Scholar] [CrossRef] [Green Version]
  89. Takakusaki, K. Neurophysiology of gait: From the spinal cord to the frontal lobe. Mov. Disord. 2013, 28, 1483–1491. [Google Scholar] [CrossRef]
  90. Faller, K.M.; Gutierrez-Quintana, R.; Mohammed, A.; Rahim, A.A.; Tuxworth, R.I.; Wager, K.; Bond, M. The neuronal ceroid lipofuscinoses: Opportunities from model systems. Biochim Biophys Acta-Mol. Basis Dis. 2015, 1852 Pt B, 2267–2278. [Google Scholar] [CrossRef] [Green Version]
  91. Kann, O. The interneuron energy hypothesis: Implications for brain disease. Neurobiol. Dis. 2016, 90, 75–85. [Google Scholar] [CrossRef]
  92. Walkley, S.U.; Sikora, J.; Micsenyi, M.; Davidson, C.; Dobrenis, K. Lysosomal compromise and brain dysfunction: Examining the role of neuroaxonal dystrophy. Biochem. Soc. Trans. 2010, 38, 1436–1441. [Google Scholar] [CrossRef]
  93. Wei, D.; Yang, F.; Wang, Y.; Yang, F.; Wu, C.; Wu, S.X.; Jiang, W. Degeneration and regeneration of GABAergic interneurons in the dentate gyrus of adult mice in experimental models of epilepsy. CNS Neurosci. Ther. 2015, 21, 52–60. [Google Scholar] [CrossRef] [PubMed]
  94. Vargova, G.; Vogels, T.; Kostecka, Z.; Hromadka, T. Inhibitory interneurons in Alzheimer’s disease. Bratisl. Lek. Listy 2018, 119, 205–209. [Google Scholar] [CrossRef] [PubMed]
  95. Tsuiji, H.; Inoue, I.; Takeuchi, M.; Furuya, A.; Yamakage, Y.; Watanabe, S.; Koike, M.; Hattori, M.; Yamanaka, K. TDP-43 accelerates age-dependent degeneration of interneurons. Sci. Rep. 2017, 7, 14972. [Google Scholar] [CrossRef] [Green Version]
  96. Conti, M.M.; Chambers, N.; Bishop, C. A new outlook on cholinergic interneurons in Parkinson’s disease and L-DOPA-induced dyskinesia. Neurosci. Biobehav. Rev. 2018, 92, 67–82. [Google Scholar] [CrossRef]
  97. Wirths, O.; Breyhan, H.; Marcello, A.; Cotel, M.C.; Brück, W.; Bayer, T.A. Inflammatory changes are tightly associated with neurodegeneration in the brain and spinal cord of the APP/PS1KI mouse model of Alzheimer’s disease. Neurobiol. Aging 2010, 31, 747–757. [Google Scholar] [CrossRef] [PubMed]
  98. Palmer, D.N.; Barry, L.A.; Tyynelä, J.; Cooper, J.D. NCL disease mechanisms. Biochim Biophys Acta-Mol. Basis Dis. 2013, 1832, 1882–1893. [Google Scholar] [CrossRef]
  99. Bosch, M.E.; Kielian, T. Neuroinflammatory paradigms in lysosomal storage diseases. Front. Neurosci. 2015, 9, 417. [Google Scholar] [CrossRef] [Green Version]
  100. Leyns, C.E.G.; Holtzman, D.M. Glial contributions to neurodegeneration in tauopathies. Mol. Neurodegener. 2017, 12, 50. [Google Scholar] [CrossRef]
  101. Wolf, S.A.; Boddeke, H.W.; Kettenmann, H. Microglia in Physiology and Disease. Annu. Rev. Physiol. 2017, 79, 619–643. [Google Scholar] [CrossRef]
  102. Groh, J.; Kühl, T.G.; Ip, C.W.; Nelvagal, H.R.; Sri, S.; Duckett, S.; Mirza, M.; Langmann, T.; Cooper, J.D.; Martini, R. Immune cells perturb axons and impair neuronal survival in a mouse model of infantile neuronal ceroid lipofuscinosis. Brain 2013, 136 Pt 4, 1083–1101. [Google Scholar] [CrossRef] [Green Version]
  103. Groh, J.; Ribechini, E.; Stadler, D.; Schilling, T.; Lutz, M.B.; Martini, R. Sialoadhesin promotes neuroinflammation-related disease progression in two mouse models of CLN disease. Glia 2016, 64, 792–809. [Google Scholar] [CrossRef] [PubMed]
  104. Seehafer, S.S.; Ramirez-Montealegre, D.; Wong, A.M.; Chan, C.H.; Castaneda, J.; Horak, M.; Ahmadi, S.M.; Lim, M.J.; Cooper, J.D.; Pearce, D.A. Immunosuppression alters disease severity in juvenile Batten disease mice. J. Neuroimmunol. 2011, 230, 169–172. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  105. Lie, P.P.Y.; Nixon, R.A. Lysosome trafficking and signaling in health and neurodegenerative diseases. Neurobiol. Dis. 2019, 122, 94–105. [Google Scholar] [CrossRef] [PubMed]
  106. Lim, C.Y.; Zoncu, R. The lysosome as a command-and-control center for cellular metabolism. J. Cell Biol. 2016, 214, 653–664. [Google Scholar] [CrossRef]
  107. Adams, J.; Feuerborn, M.; Molina, J.A.; Wilden, A.R.; Adhikari, B.; Budden, T.; Lee, S.Y. Autophagy-lysosome pathway alterations and alpha-synuclein up-regulation in the subtype of neuronal ceroid lipofuscinosis, CLN5 disease. Sci. Rep. 2019, 9, 151. [Google Scholar] [CrossRef] [Green Version]
  108. Best, H.L.; Neverman, N.J.; Wicky, H.E.; Mitchell, N.L.; Leitch, B.; Hughes, S.M. Characterisation of early changes in ovine CLN5 and CLN6 Batten disease neural cultures for the rapid screening of therapeutics. Neurobiol. Dis. 2017, 100, 62–74. [Google Scholar] [CrossRef] [PubMed]
  109. Ward, C.; Martinez-Lopez, N.; Otten, E.G.; Carroll, B.; Maetzel, D.; Singh, R.; Sarkar, S.; Korolchuk, V.I. Autophagy, lipophagy and lysosomal lipid storage disorders. Biochim. Biophys. Acta-Mol. Cell Biol. Lipids 2016, 1861, 269–284. [Google Scholar] [CrossRef] [Green Version]
  110. Lieberman, A.P.; Puertollano, R.; Raben, N.; Slaugenhaupt, S.; Walkley, S.U.; Ballabio, A. Autophagy in lysosomal storage disorders. Autophagy 2012, 8, 719–730. [Google Scholar] [CrossRef] [Green Version]
  111. Boya, P. Lysosomal function and dysfunction: Mechanism and disease. Antioxid. Redox Signal. 2012, 17, 766–774. [Google Scholar] [CrossRef] [Green Version]
  112. Llavero Hurtado, M.; Fuller, H.R.; Wong, A.; Eaton, S.L.; Gillingwater, T.H.; Pennetta, G.; Cooper, J.D.; Wishart, T.M. Proteomic mapping of differentially vulnerable pre-synaptic populations identifies regulators of neuronal stability in vivo. Sci. Rep. 2017, 7, 12412. [Google Scholar] [CrossRef] [Green Version]
  113. Koster, K.P.; Francesconi, W.; Berton, F.; Alahmadi, S.; Srinivas, R.; Yoshii, A. Developmental NMDA receptor dysregulation in the infantile neuronal ceroid lipofuscinosis mouse model. eLife 2019, 8, e40316. [Google Scholar] [CrossRef] [PubMed]
  114. Lehtovirta, M.; Kyttälä, A.; Eskelinen, E.L.; Hess, M.; Heinonen, O.; Jalanko, A. Palmitoyl protein thioesterase (PPT) localizes into synaptosomes and synaptic vesicles in neurons: Implications for infantile neuronal ceroid lipofuscinosis (INCL). Hum. Mol. Genet. 2001, 10, 69–75. [Google Scholar] [CrossRef] [PubMed]
  115. Di Giovanni, J.; Sheng, Z.H. Regulation of synaptic activity by snapin-mediated endolysosomal transport and sorting. EMBO J. 2015, 34, 2059–2077. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  116. Goo, M.S.; Sancho, L.; Slepak, N.; Boassa, D.; Deerinck, T.J.; Ellisman, M.H.; Bloodgood, B.L.; Patrick, G.N. Activity-dependent trafficking of lysosomes in dendrites and dendritic spines. J. Cell Biol. 2017, 216, 2499–2513. [Google Scholar] [CrossRef] [Green Version]
  117. Okerlund, N.D.; Schneider, K.; Leal-Ortiz, S.; Montenegro-Venegas, C.; Kim, S.A.; Garner, L.C.; Waites, C.L.; Gundelfinger, E.D.; Reimer, R.J.; Garner, C.C. Bassoon Controls Presynaptic Autophagy through Atg5. Neuron 2017, 93, 897–913. [Google Scholar] [CrossRef] [Green Version]
  118. Li, Y.C.; Kavalali, E.T. Synaptic Vesicle-Recycling Machinery Components as Potential Therapeutic Targets. Pharmacol. Rev. 2017, 69, 141–160. [Google Scholar] [CrossRef] [Green Version]
  119. Blanco-Suárez, E.; Caldwell, A.L.; Allen, N.J. Role of astrocyte-synapse interactions in CNS disorders. J. Physiol. 2017, 595, 1903–1916. [Google Scholar] [CrossRef] [Green Version]
  120. Colonna, M.; Butovsky, O. Microglia Function in the Central Nervous System During Health and Neurodegeneration. Annu. Rev. Immunol. 2017, 35, 441–468. [Google Scholar] [CrossRef]
  121. Singh, A.; Abraham, W.C. Astrocytes and synaptic plasticity in health and disease. Exp. Brain Res. 2017, 235, 1645–1655. [Google Scholar] [CrossRef]
  122. Settembre, C.; Fraldi, A.; Medina, D.L.; Ballabio, A. Signals from the lysosome: A control centre for cellular clearance and energy metabolism. Nat. Rev. Mol. Cell Biol. 2013, 14, 283–296. [Google Scholar] [CrossRef] [Green Version]
  123. Fraldi, A.; Klein, A.D.; Medina, D.L.; Settembre, C. Brain Disorders Due to Lysosomal Dysfunction. Annu. Rev. Neurosci. 2016, 39, 277–295. [Google Scholar] [CrossRef] [PubMed]
  124. Sharma, J.; di Ronza, A.; Lotfi, P.; Sardiello, M. Lysosomes and Brain Health. Annu. Rev. Neurosci. 2018, 41, 255–276. [Google Scholar] [CrossRef] [PubMed]
  125. Bagh, M.B.; Peng, S.; Chandra, G.; Zhang, Z.; Singh, S.P.; Pattabiraman, N.; Liu, A.; Mukherjee, A.B. Misrouting of v-ATPase subunit V0a1 dysregulates lysosomal acidification in a neurodegenerative lysosomal storage disease model. Nat. Commun. 2017, 8, 14612. [Google Scholar] [CrossRef] [Green Version]
  126. Danyukova, T.; Ariunbat, K.; Thelen, M.; Brocke-Ahmadinejad, N.; Mole, S.E.; Storch, S. Loss of CLN7 results in depletion of soluble lysosomal proteins and impaired mTOR reactivation. Hum. Mol. Genet. 2018, 27, 1711–1722. [Google Scholar] [CrossRef] [PubMed]
  127. Holler, C.J.; Taylor, G.; McEachin, Z.T.; Deng, Q.; Watkins, W.J.; Hudson, K.; Easley, C.A.; Hu, W.T.; Hales, C.M.; Rossoll, W.; et al. Trehalose upregulates progranulin expression in human and mouse models of GRN haploinsufficiency: A novel therapeutic lead to treat frontotemporal dementia. Mol. Neurodegener. 2016, 11, 46. [Google Scholar] [CrossRef] [Green Version]
  128. Palmieri, M.; Pal, R.; Nelvagal, H.R.; Lotfi, P.; Stinnett, G.R.; Seymour, M.L.; Chaudhury, A.; Bajaj, L.; Bondar, V.V.; Bremner, L.; et al. mTORC1-independent TFEB activation via Akt inhibition promotes cellular clearance in neurodegenerative storage diseases. Nat. Commun. 2017, 8, 14338. [Google Scholar] [CrossRef]
  129. Mole, S.E.; Cotman, S.L. Genetics of the neuronal ceroid lipofuscinoses (Batten disease). Biochim. Biophys. Acta-Mol. Basis Dis. 2015, 1852 Pt B, 2237–2241. [Google Scholar] [CrossRef] [Green Version]
  130. Huber, R.J.; Mathavarajah, S. Cln5 is secreted and functions as a glycoside hydrolase in Dictyostelium. Cell. Signal. 2018, 42, 236–248. [Google Scholar] [CrossRef]
  131. Adams, H.R.; Mink, J.W. Neurobehavioral features and natural history of juvenile neuronal ceroid lipofuscinosis (Batten disease). J. Child. Neurol. 2013, 28, 1128–1136. [Google Scholar] [CrossRef] [Green Version]
  132. Steinfeld, R.; Heim, P.; von Gregory, H.; Meyer, K.; Ullrich, K.; Goebel, H.H.; Kohlschütter, A. Late infantile neuronal ceroid lipofuscinosis: Quantitative description of the clinical course in patients with CLN2 mutations. Am. J. Med. Genet. 2002, 112, 347–354. [Google Scholar] [CrossRef]
  133. de Blieck, E.A.; Augustine, E.F.; Marshall, F.J.; Adams, H.; Cialone, J.; Dure, L.; Kwon, J.M.; Newhouse, N.; Rose, K.; Rothberg, P.G.; et al. Methodology of clinical research in rare diseases: Development of a research program in juvenile neuronal ceroid lipofuscinosis (JNCL) via creation of a patient registry and collaboration with patient advocates. Contemp. Clin. Trials. 2013, 35, 48–54. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  134. Nickel, M.; Simonati, A.; Jacoby, D.; Lezius, S.; Kilian, D.; Van de Graaf, B.; Pagovich, O.E.; Kosofsky, B.; Yohay, K.; Downs, M.; et al. Disease characteristics and progression in patients with late-infantile neuronal ceroid lipofuscinosis type 2 (CLN2) disease: An observational cohort study. Lancet Child. Adolesc. Health 2018, 2, 582–590. [Google Scholar] [CrossRef]
  135. Simonati, A.; Williams, R.E.; Nardocci, N.; Laine, M.; Battini, R.; Schulz, A.; Garavaglia, B.; Moro, F.; Pezzini, F.; Santorelli, F.M. Phenotype and natural history of variant late infantile ceroid-lipofuscinosis 5. Dev. Med. Child. Neurol. 2017, 59, 815–821. [Google Scholar] [CrossRef] [Green Version]
  136. Schulz, A.; Ajayi, T.; Specchio, N.; de Los Reyes, E.; Gissen, P.; Ballon, D.; Dyke, J.P.; Cahan, H.; Slasor, P.; Jacoby, D.; et al. Study of Intraventricular Cerliponase Alfa for CLN2 Disease. N. Engl. J. Med. 2018, 378, 1898–1907. [Google Scholar] [CrossRef]
  137. Canafoglia, L.; Gilioli, I.; Invernizzi, F.; Sofia, V.; Fugnanesi, V.; Morbin, M.; Chiapparini, L.; Granata, T.; Binelli, S.; Scaioli, V.; et al. Electroclinical spectrum of the neuronal ceroid lipofuscinoses associated with CLN6 mutations. Neurology 2015, 85, 316–324. [Google Scholar] [CrossRef] [Green Version]
  138. Anderson, G.; Smith, V.V.; Malone, M.; Sebire, N.J. Blood film examination for vacuolated lymphocytes in the diagnosis of metabolic disorders; retrospective experience of more than 2500 cases from a single centre. J. Clin. Pathol. 2005, 58, 1305–1310. [Google Scholar] [CrossRef] [Green Version]
  139. Simonati, A.; Williams, R.E. Neuronal Ceroid Lipofuscinosis: The Multifaceted Approach to the Clinical Issues, an Overview. Front. Neurol. 2022, 13, 811686. [Google Scholar] [CrossRef]
  140. Dilaveris, P.; Koutagiar, I.; Aggeli, C.; Sideris, S.; Gatzoulis, K.; Stefanadis, C. Severe sinus node dysfunction in a patient with juvenile neuronal ceroid lipofuscinosis. Int. J. Cardiol. 2014, 174, 143–146. [Google Scholar] [CrossRef] [PubMed]
  141. Williams, R.E.; Adams, H.R.; Blohm, M.; Cohen-Pfeffer, J.L.; de Los Reyes, E.; Denecke, J.; Drago, K.; Fairhurst, C.; Frazier, M.; Guelbert, N.; et al. Management Strategies for CLN2 Disease. Pediatr. Neurol. 2017, 69, 102–112. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  142. Augustine, E.F.; Adams, H.R.; Beck, C.A.; Vierhile, A.; Kwon, J.; Rothberg, P.G.; Marshall, F.; Block, R.; Dolan, J.; Mink, J.W.; et al. Standardized assessment of seizures in patients with juvenile neuronal ceroid lipofuscinosis. Dev. Med. Child. Neurol. 2015, 57, 366–371. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  143. Johannsen, J.; Nickel, M.; Schulz, A.; Denecke, J. Considering Valproate as a Risk Factor for Rapid Exacerbation of Complex Movement Disorder in Progressed Stages of Late-Infantile CLN2 Disease. Neuropediatrics 2016, 47, 194–196. [Google Scholar] [CrossRef] [PubMed]
  144. Fietz, M.; AlSayed, M.; Burke, D.; Cohen-Pfeffer, J.; Cooper, J.D.; Dvořáková, L.; Giugliani, R.; Izzo, E.; Jahnová, H.; Lukacs, Z.; et al. Diagnosis of neuronal ceroid lipofuscinosis type 2 (CLN2 disease): Expert recommendations for early detection and laboratory diagnosis. Mol. Genet. Metab. 2016, 119, 160–167. [Google Scholar] [CrossRef] [Green Version]
  145. Aldrich, A.; Kielian, T. Central nervous system fibrosis is associated with fibrocyte-like infiltrates. Am. J. Pathol. 2011, 179, 2952–2962. [Google Scholar] [CrossRef] [PubMed]
  146. Kohan, R.; Cismondi, I.A.; Kremer, R.D.; Muller, V.J.; Guelbert, N.; Anzolini, V.T.; Fietz, M.J.; Ramírez, A.M.; Halac, I.N. An integrated strategy for the diagnosis of neuronal ceroid lipofuscinosis types 1 (CLN1) and 2 (CLN2) in eleven Latin American patients. Clin. Genet. 2009, 76, 372–382. [Google Scholar] [CrossRef] [PubMed]
  147. Kohan, R.; Pesaola, F.; Guelbert, N.; Pons, P.; Oller-Ramírez, A.M.; Rautenberg, G.; Becerra, A.; Sims, K.; Xin, W.; Cismondi, I.A.; et al. The neuronal ceroid lipofuscinoses program: A translational research experience in Argentina. Biochim. Biophys. Acta-Mol. Basis Dis. 2015, 1852 Pt B, 2301–2311. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  148. Patiño, L.C.; Battu, R.; Ortega-Recalde, O.; Nallathambi, J.; Anandula, V.R.; Renukaradhya, U.; Laissue, P. Exome sequencing is an efficient tool for variant late-infantile neuronal ceroid lipofuscinosis molecular diagnosis. PLoS ONE 2014, 9, e109576. [Google Scholar] [CrossRef]
  149. Fritchie, K.; Siintola, E.; Armao, D.; Lehesjoki, A.E.; Marino, T.; Powell, C.; Tennison, M.; Booker, J.M.; Koch, S.; Partanen, S.; et al. Novel mutation and the first prenatal screening of cathepsin D deficiency (CLN10). Acta Neuropathol. 2009, 117, 201–208. [Google Scholar] [CrossRef]
  150. Mole, S.E.; Williams, R.E. Neuronal ceroid-lipofuscinoses. In GeneReviews. Available online: https://0-www-ncbi-nlm-nih-gov.brum.beds.ac.uk/books/NBK1428 (accessed on 16 May 2022).
  151. Junaid, M.A.; Sklower, B.S.; Wisniewski, K.E.; Pullarkat, R.K. A novel assay for lysosomal pepstatin-insensitive proteinase and its application for the diagnosis of late-infantile neuronal ceroid lipofuscinosis. Clin. Chim. Acta 1999, 281, 169–176. [Google Scholar] [CrossRef]
  152. van Diggelen, O.P.; Keulemans, J.L.; Winchester, B.; Hofman, I.L.; Vanhanen, S.L.; Santavuori, P.; Voznyi, Y.V. A rapid fluorogenic palmitoyl-protein thioesterase assay: Pre-and postnatal diagnosis of INCL. Mol. Genet. Metab. 1999, 66, 240–244. [Google Scholar] [CrossRef]
  153. Voznyi, Y.V.; Keulemans, J.L.; Mancini, G.M.; Catsman-Berrevoets, C.E.; Young, E.; Winchester, B.; Kleijer, W.J.; van Diggelen, O.P. A new simple enzyme assay for pre-and postnatal diagnosis of infantile neuronal ceroid lipofuscinosis (INCL) and its variants. J. Med. Genet. 1999, 36, 471–474. [Google Scholar]
  154. Vines, D.J.; Warburton, M.J. Classical late infantile neuronal ceroid lipofuscinosis fibroblasts are deficient in lysosomal tripeptidyl peptidase I. FEBS Lett. 1999, 443, 131–135. [Google Scholar] [CrossRef] [Green Version]
  155. Sohar, I.; Lin, L.; Lobel, P. Enzyme-based diagnosis of classical late infantile neuronal ceroid lipofuscinosis: Comparison of tripeptidyl peptidase I and pepstatin-insensitive protease assays. Clin. Chem. 2000, 46, 1005–1008. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  156. Partanen, S.; Storch, S.; Löffler, H.G.; Hasilik, A.; Tyynelä, J.; Braulke, T. A replacement of the active-site aspartic acid residue 293 in mouse cathepsin D affects its intracellular stability, processing and transport in HEK-293 cells. Biochem. J. 2003, 369 Pt 1, 55–62. [Google Scholar] [CrossRef] [PubMed]
  157. Ceuterick-de Groote, C.; Martin, J.J. Extracerebral biopsy in lysosomal and peroxisomal disorders. Ultrastructural findings. Brain Pathol. 1998, 8, 121–132. [Google Scholar] [CrossRef] [PubMed]
  158. Williams, R.E.; Aberg, L.; Autti, T.; Goebel, H.H.; Kohlschütter, A.; Lönnqvist, T. Diagnosis of the neuronal ceroid lipofuscinoses: An update. Biochim. Biophys. Acta-Mol. Basis Dis. 2006, 1762, 865–872. [Google Scholar] [CrossRef] [Green Version]
  159. Dyke, J.P.; Sondhi, D.; Voss, H.U.; Yohay, K.; Hollmann, C.; Mancenido, D.; Kaminsky, S.M.; Heier, L.A.; Rudser, K.D.; Kosofsky, B.; et al. Brain Region-Specific Degeneration with Disease Progression in Late Infantile Neuronal Ceroid Lipofuscinosis (CLN2 Disease). Am. J. Neuroradiol. 2016, 37, 1160–1169. [Google Scholar] [CrossRef] [Green Version]
  160. Specchio, N.; Bellusci, M.; Pietrafusa, N.; Trivisano, M.; de Palma, L.; Vigevano, F. Photosensitivity is an early marker of neuronal ceroid lipofuscinosis type 2 disease. Epilepsia 2017, 58, 1380–1388. [Google Scholar] [CrossRef] [Green Version]
  161. Albert, D.V.; Yin, H.; De Los Reyes, E.C.; Vidaurre, J. Unique Characteristics of the Photoparoxysmal Response in Patients with Neuronal Ceroid Lipofuscinosis Type 2: Can EEG Be a Biomarker? J. Child. Neurol. 2016, 31, 1475–1482, Correction in J. Child. Neurol. 2018, 33, 748. [Google Scholar] [CrossRef]
  162. Özkara, Ç.; Gündüz, A.; Coşkun, T.; Alpaslan, B.G.; Zeydan, B.; Delil, Ş.; Muona, M.; Lehesjoki, A.E.; Kızıltan, M.E. Long-term follow-up of two siblings with adult-onset neuronal ceroid lipofuscinosis, Kufs type A. Epileptic Disord. 2017, 19, 147–151. [Google Scholar] [CrossRef]
  163. Lv, Y.; Zhang, N.; Liu, C.; Shi, M.; Sun, L. Occipital epilepsy versus progressive myoclonic epilepsy in a patient with continuous occipital spikes and photosensitivity in electroencephalogram: A case report. Medicine 2018, 97, e0299. [Google Scholar] [CrossRef]
  164. Pampiglione, G.; Harden, A. So-called neuronal ceroid lipofuscinosis. J. Neurol. Neurosurg. Psychiatry 1977, 40, 323–330. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  165. Trivisano, M.; Ferretti, A.; Calabrese, C.; Pietrafusa, N.; Piscitello, L.; Carfi’ Pavia, G.; Vigevano, F.; Specchio, N. Neurophysiological Findings in Neuronal Ceroid Lipofuscinoses. Front. Neurol. 2022, 13, 845877. [Google Scholar] [CrossRef] [PubMed]
  166. Dozières-Puyravel, B.; Nasser, H.; Elmaleh-Bergès, M.; Lopez Hernandez, E.; Gelot, A.; Ilea, A.; Delanoë, C.; Puech, J.P.; Caillaud, C.; Pichard, S.; et al. Paediatric-onset neuronal ceroid lipofuscinosis: First symptoms and presentation at diagnosis. Dev. Med. Child. Neurol. 2020, 62, 528–530. [Google Scholar] [CrossRef] [PubMed]
  167. Beltrán, L.; Valenzuela, G.R.; Loos, M.; Vargas, R.; Lizama, R.; Spinsanti, P.; Caraballo, R. Late-onset childhood neuronal ceroid lipofuscinosis: Early clinical and electroencephalographic markers. Epilepsy Res. 2018, 144, 49–52. [Google Scholar] [CrossRef]
  168. Kasteleijn-Nolst Trenité, D.; Mastrangelo, M.; Spalice, A.; Parisi, P. Photosensitivity as an early marker of epileptic and developmental encephalopathies. Epilepsia 2018, 59, 1086–1087. [Google Scholar] [CrossRef] [Green Version]
  169. Thompson, D.A.; Handley, S.E.; Henderson, R.H.; Marmoy, O.R.; Gissen, P. An ERG and OCT study of neuronal ceroid lipofuscinosis CLN2 Battens retinopathy. Eye 2021, 35, 2438–2448. [Google Scholar] [CrossRef]
  170. Simonati, A.; Tessa, A.; Bernardina, B.D.; Biancheri, R.; Veneselli, E.; Tozzi, G.; Bonsignore, M.; Grosso, S.; Piemonte, F.; Santorelli, F.M. Variant late infantile neuronal ceroid lipofuscinosis because of CLN1 mutations. Pediatr. Neurol. 2009, 40, 271–276. [Google Scholar] [CrossRef]
  171. Holmberg, V.; Lauronen, L.; Autti, T.; Santavuori, P.; Savukoski, M.; Uvebrant, P.; Hofman, I.; Peltonen, L.; Järvelä, I. Phenotype-genotype correlation in eight patients with Finnish variant late infantile NCL (CLN5). Neurology 2000, 55, 579–581. [Google Scholar] [CrossRef]
  172. Jadav, R.H.; Sinha, S.; Yasha, T.C.; Aravinda, H.; Gayathri, N.; Rao, S.; Bindu, P.S.; Satishchandra, P. Clinical, electrophysiological, imaging, and ultrastructural description in 68 patients with neuronal ceroid lipofuscinoses and its subtypes. Pediatr. Neurol. 2014, 50, 85–95. [Google Scholar] [CrossRef]
  173. Quagliato, E.M.A.B.; Rocha, D.M.; Sacai, P.Y.; Watanabe, S.S.; Salom~ao, S.R.; Berezovsky, A. Retinal function in patients with the neuronal ceroid lipofuscinosis phenotype. Arq. Bras. Oftalmol. 2017, 80, 215–219. [Google Scholar] [CrossRef]
  174. Shen, J.; Cram, D.S.; Wu, W.; Cai, L.; Yang, X.; Sun, X.; Cui, Y.; Liu, J. Successful PGD for late infantile neuronal ceroid lipofuscinosis achieved by combined chromosome and TPP1 gene analysis. Reprod. Biomed. Online 2013, 27, 176–183. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  175. Goebel, H.H.; Vesa, J.; Reitter, B.; Goecke, T.O.; Schneider-Rätzke, B.; Merz, E. Prenatal diagnosis of infantile neuronal ceroid-lipofuscinosis: A combined electron microscopic and molecular genetic approach. Brain Dev. 1995, 17, 83–88. [Google Scholar] [CrossRef]
  176. Lake, B.D.; Young, E.P.; Winchester, B.G. Prenatal diagnosis of lysosomal storage diseases. Brain Pathol. 1998, 8, 133–149. [Google Scholar] [CrossRef] [PubMed]
  177. Munroe, P.B.; Rapola, J.; Mitchison, H.M.; Mustonen, A.; Mole, S.E.; Gardiner, R.M.; Jarvela, I. Prenatal diagnosis of Batten’s disease. Lancet 1996, 347, 1014–1015. [Google Scholar] [CrossRef]
Table 1. Types of NCL.
Table 1. Types of NCL.
ACongenital NCL
Babies are born with microcephaly as the disease begins in utero [21,27].
BInfantile NCL
Seizures and the loss of motor function appear between the ages of 6 and 18 months, with the loss of psychomotor skills, including speech.
Signs of regression, accompanied by the onset of epilepsy and a progressive loss of vision.
Hyperexcitability.
Restlessness and poor sleep.
After the age of 15 to 20 months, the acceleration of symptoms occurs, leading to microcephaly, truncal ataxia, dystonic features, choreoathetosis, and myoclonic jerks.
By the age of 24 months, children become blind and lose all cognitive and active motor skill.
Death usually between the ages of 9 and 13 [21,28].
CLate infantile NCL
Developmental delay, ataxia, and seizures appear between the ages of 2 and 4 years old and progress rapidly to loss of motor, cognitive, and language functions, ultimately becoming behaviorally abnormal and demented [21,28].
DJuvenile NCL (JNCL)
The most common type of NCL.
Symptoms occurring between 5 and 10 years old—commonly associated with the progressive loss of vision and seizures.
Learning difficulties and motor disturbances, including extrapyramidal and less prominent pyramidal involvement (rigidity, bradykinesia, slow steps with flexion in hips and knees, and shuffling gait), appear around puberty.
Death during their third decade [21,28].
EAdult NCL
Symptoms are less severe and progress more slowly.
The clinical picture is characterized by generalized tonic seizures, myoclonus, and prominent dementia.
Associated features include speech problems, cerebellar dysfunction, and parkinsonism [21,29,30].
Table 2. NCL diseases and encoded genes [19,21,31,32].
Table 2. NCL diseases and encoded genes [19,21,31,32].
DiseaseGene/ProteinAge of OnsetProtein Function
CLN1PPT1 (palmitoyl protein thioesterase 1)6–18 monthsPalmitoyl-protein thioesterase activity plays a critical role in the degradation of lipid-modified proteins via removing fatty acid residues from cysteine residues.
CLN2TPP1 (tripeptidyl peptidase 1)2–4 yearsSerine protease activity prevents intralysosomal accumulation of storage material and neuronal loss.
CLN3CLN3, lysosomal/endosomal transmembrane protein4–10 yearsPredicted function as a pH regulator and modulator of vesicular trafficking and fusion that promotes cellular homeostasis and neuronal survival.
CLN4DNAJC5/CSPα (cysteine string protein α)>18 yearsInvolvement in exocytosis and endocytosis functions plays a regulatory role in ATPase activity and assists in folding proteins in synaptic vesicles.
CLN5Soluble lysosomal protein4–7 yearsGlycoside hydrolase activity modulates vesicular trafficking.
CLN6Transmembrane protein of endoplasmic reticulum18 months–6 yearsPrecise function remains unclear but is linked with intracellular trafficking and lysosomal function.
CLN7MFSD8 (major facilitator superfamily domain-containing 8), lysosomal transmembrane protein2–6 yearsPredicted transmembrane transporter function plays a role in preventing neuronal loss, robust accumulation of lipofuscin, reactive gliosis, and degeneration and storage accumulation in the retina.
CLN8Transmembrane protein of endoplasmic reticulum2–7 years (Turkish variant late-infantile NCL)Aids in lysosomal biogenesis, through transportation from the ER to the Golgi complex, as well as in the regulation of lipid metabolism.
5–10 years (northern epilepsy)
CLN10CTSD (cathepsin D)In uteroAspartic protease functions in an unknown neuroprotective mechanism.
CLN11PRGN (progranulin)20–25 yearsKnown roles in inflammation, embryogenesis, cell motility, and tumorigenesis.
CLN12ATP13A213–16 yearsRegulation of ion homeostasis.
CLN13CTSF (cathepsin F)>18 yearsLoss of lysosomal cysteine protease activity leads to the deterioration of motor function and reduced brain function.
CLN14KCTD7 (potassium channel tetramerization domain-containing protein 7)8–24 monthsModulation of potassium ion channel activity.
Table 3. Major clinical manifestations [3,4,13].
Table 3. Major clinical manifestations [3,4,13].
DiseaseClinical Manifestation
CLN1Cognitive and motor decline
Ataxia
Choreoathetosis
Myoclonus, epilepsy
Progressive loss of vision
Decelerated head growth
Irritability
Hyperexcitability
Death by 10 years
CLN2Cognitive decline
Motor coordination loss
Ataxia
Choreoathetosis
Myoclonus, spasticity
Dystonic movements
Epilepsy
Progressive loss of vision
Behavioural disturbances
Death in early adolescence
CLN3Progressive loss of vision
Pigmentary retinopathy
Motor e cognitive decline
Epilepsy
Behavioural disturbances
Spasticity, myoclonus
Death in the second decade
CLN4Cognitive and motor decline
Epilepsy, myoclonus
Ataxia, tremors
Behavioural disturbances
No visual impairment
Death within 15 years after onset
CLN5Psychomotor regression
Ataxia
Motor coordination loss
Epilepsy
Progressive loss of vision
Macular degeneration
Death between 13–30 years
CLN6Motor skills loss
Dysarthria
Ataxia
Epilepsy
Progressive loss of vision
Death in the second decade
CLN7Cognitive and motor decline
Ataxia, myoclonus
Epilepsy
Progressive loss of vision
Behavioural disturbances
Death in the second decade
CLN8Cognitive and motor decline
Ataxia
Myoclonus
Epilepsy
Progressive loss of vision
Retinopathy
Behavioural disturbances
Death within the sixth decade
CLN10Microencephaly
Hypotonia
Absence of reflexes
Respiratory impairment at birth
Epilepsy
Death within hours after birth
CLN11Rapid progressive retinal dystrophy and progressive loss of vision
Cognitive decline
Epilepsy
Myoclonus
Ataxia
CLN12Learning disorders
Loss of speech
Loss of coordination
Ataxia, myoclonus
Epilepsy
Spasticity
Behavioural disturbances
Muscular atrophy
No visual impairment
CLN13Memory deficits
Behavioural disturbances
Tremors, bradykinesia
Ataxia, rigidity
Dysarthria
Epilepsy
No visual impairment
CLN14Motor decline
Epilepsy
Myoclonus
Ataxia
Dysarthria
Optic atrophy
Progressive loss of vision
Behavioural disturbances
Table 4. NCL subtypes and abnormalities in diagnostic test results [3,4,16,20].
Table 4. NCL subtypes and abnormalities in diagnostic test results [3,4,16,20].
DiseaseNeuroradiological Imaging FindingsEEG AbnormalitiesVisual ChangesMicroscopic Findings
CLN1Hyperintense, periventricular high-signal rims of white matter; decreased NAA and increased choline; and severely enlarged lateral ventriclesLoss of sleep spindles at ~2 years; attenuated reaction to passive eye opening and/or closing; background activity disturbances; and reduced amplitudeOptic atrophy; unrecordable ERG at 4 years; and blindnessGRODs
CLN2Infratentorial atrophy; hypointense thalamic nuclei; decreased NAA; increased myoinositol and Glu:Gln ratio in the white matter; and severely enlarged lateral ventriclesOccipital spike in response to slow flash; irregular slow activity; focal spikes; and absence of sleep spindlesProgressive loss of vision and diminished ERGCLPs
CLN3Cerebellar atrophy; enlarged third ventricle and cerebral sulci; and hypointense thalamic nucleiProgressive background disorganization; and spike-and-slow-wave complexesProgressive loss of vision and pigmentary retinopathyFPPs and vacuolated lymphocytes
CLN4Parieto-occipital cortical atrophy; cerebellar atrophy; hyperintense periventricular areas; and corpus callosum thinningSlow background; polyphasic spikes; and slow-wave dischargesNo visual impairmentGRODs; CLPs; and FPPs
CLN5Cerebellar atrophy; diminished signal intensity in thalamic nuclei; and increased signal intensity in periventricular white matter and internal capsuleOccipital spikes in response to slow flashProgressive loss of vision and macular degenerationGRODs; CLPs; and FPPs
CLN6Deep cortical layer-specific neuron loss and cerebellar atrophyBackground slowing; and high-amplitude discharges, in response to photic stimulationProgressive loss of visionCLPs; RLC; and FPPs
CLN7Cerebellar atrophy; corpus callosum thinning; and hypointense thalamic nucleiOccipital spikes; and background slowingProgressive loss of visionCLPs; RLC; and FPPs
CLN8Cerebellar atrophy; corpus callosum thinning; and hyperintensity of white matterSlow background; components of high amplitude; and epileptiform dischargesRetinopathy; visual decline at around 4–6 years of age; and ERG absentGRODs; CLPs; and FPPs
CLN10Diminished head growth in utero; myoclonic fetal seizures; enlarged lateral ventricles; hypointense cerebral and cerebellar white matter; and decreased NAA and increase in myo-inositolCompletely depleted EEG patternNDGRODs
CLN11Cerebellar atrophyPolyspike-wave discharges with posterior emphasis; and severe attenuation of rod and cone responsesProgressive loss of vision and retinal dystrophyFPPs
CLN12Cortical and subcortical atrophy; decreased glucose use in grey matter, especially the thalamus and posterior association cortexNDNo visual changesGRODs
CLN13Cerebellar atrophy; frontal and parietal cortical atrophy; and periventricular hyperintensitiesNo epileptiform activityNDFPPs
CLN14Cortical and cerebellar atrophy; and corpus callosum thinningSlow dysrhythmia; multifocal high-amplitude epileptiform discharges; photosensitivity; and occipital spikesProgressive loss of vision; diminished pupillary light reflex; and optic atrophyGRODs; CLPs; RLC; and FPPs
Abbreviations: NAA = N-acetyl aspartate; ERG = electroretinogram; GRODs = granular deposits; CLPs = urvilinear profiles; FPPs = fingerprint profiles; RLC = rectilinear complex; ND = not determined.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Kaminiów, K.; Kozak, S.; Paprocka, J. Recent Insight into the Genetic Basis, Clinical Features, and Diagnostic Methods for Neuronal Ceroid Lipofuscinosis. Int. J. Mol. Sci. 2022, 23, 5729. https://0-doi-org.brum.beds.ac.uk/10.3390/ijms23105729

AMA Style

Kaminiów K, Kozak S, Paprocka J. Recent Insight into the Genetic Basis, Clinical Features, and Diagnostic Methods for Neuronal Ceroid Lipofuscinosis. International Journal of Molecular Sciences. 2022; 23(10):5729. https://0-doi-org.brum.beds.ac.uk/10.3390/ijms23105729

Chicago/Turabian Style

Kaminiów, Konrad, Sylwia Kozak, and Justyna Paprocka. 2022. "Recent Insight into the Genetic Basis, Clinical Features, and Diagnostic Methods for Neuronal Ceroid Lipofuscinosis" International Journal of Molecular Sciences 23, no. 10: 5729. https://0-doi-org.brum.beds.ac.uk/10.3390/ijms23105729

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop