- Research
- Open access
- Published:
Common molecular and pathophysiological underpinnings of delirium and Alzheimer’s disease: molecular signatures and therapeutic indications
BMC Geriatrics volume 24, Article number: 716 (2024)
Abstract
Background
Delirium and Alzheimer's disease (AD) are common causes of cognitive dysfunction among older adults. These neurodegenerative diseases share a common and complex relationship, and can occur individually or concurrently, increasing the chance of permanent mental dysfunction. However, the common molecular pathophysiology, key proteomic biomarkers, and functional pathways are largely unknown, whereby delirium is superimposed on AD and dementia.
Methods
We employed an integrated bioinformatics and system biology analysis approach to decipher such common key proteomic signatures, pathophysiological links between delirium and AD by analyzing the gene expression data of AD-affected human brain samples and comparing them with delirium-associated proteins. The present study identified the common drug target hub-proteins examining the protein–protein interaction (PPI) and gene regulatory network analysis. The functional enrichment and pathway analysis was conducted to reveal the common pathophysiological relationship. Finally, the molecular docking and dynamic simulation was used to computationally identify and validate the potential drug target and repurposable drugs for delirium and AD.
Results
We detected 99 shared differentially expressed genes (sDEGs) associated with AD and delirium. The sDEGs-set enrichment analysis detected the transmission across chemical synapses, neurodegeneration pathways, neuroinflammation and glutamatergic signaling pathway, oxidative stress, and BDNF signaling pathway as the most significant signaling pathways shared by delirium and AD. The disease-sDEGs interaction analysis highlighted the other disease risk factors with delirium and AD development and progression. Among the sDEGs of delirium and AD, the top 10 hub-proteins including ALB, APP, BDNF, CREB1, DLG4, GAD1, GAD2, GFAP, GRIN2B and GRIN2A were found by the PPI network analysis. Based on the maximum molecular docking binding affinities and molecular dynamic simulation (100 ns) results, the ALB and GAD2 were found as prominent drug target proteins when tacrine and donepezil were identified as potential drug candidates for delirium and AD.
Conclusion
The study outlined the common key biomolecules and biological pathways shared by delirium and AD. The computationally reported potential drug molecules need a deeper investigation including clinical trials to validate their effectiveness. The outcomes from this study will help to understand the typical pathophysiological relationship between delirium and AD and flag future therapeutic development research for delirium.
Background
Delirium among older adults results in a higher economic burden for the family and caregivers and increases cognitive and physical dysfunctions [1, 2]. Due to mysterious and multifactorial properties, more than half of this neurologically complicated condition remains undiagnosed in intensive care units (ICU) [3,4,5]. This is mainly caused by inconsistency in delirium’s definition and subsequent identification, and its complex molecular pathophysiology [6]. Delirium is triggered by multiple potential factors and causes, including the predisposing of older persons and potential frailty [7], pre-diagnosed cognitive impairment [8, 9], any psychological illness [10, 11], use of alcohol, and associated malnutrition [12, 13] other precipitating factors including chronic and acute medical conditions, severe diseases, trauma, major surgery and stress, and medications [4, 14]. The molecular pathophysiological mechanism of delirium involves different important signaling pathways and biological mechanisms. The oxidative stress-associated medical condition, hypoxia, is also considered a driver of delirium [15, 16]. Severe systemic cytokines-derived inflammation and peripheral neuroinflammation are widely reported and described as influencers and triggers of delirium [17, 18]. The functional disruption of brain neurotransmitter systems, including the dopamine, acetylcholine (ACh), and GABA associate pathways, and cholinergic synapses neurodegeneration, are closely connected with delirium [19, 20].
AD is an elusive neurodegenerative disease characterized by chronic and persistent cognitive impairment and is considered one of the major causes of dementia [21]. The molecular pathophysiology of AD reveals a wide range of neurobiological functions, including amyloid plaques, neurofibrillary tangles, neuroinflammation, oxidative stress, and damage to cholinergic neurons [22, 23]. The chemical synapse-associated pathways, notably axonal dystrophy, loss of pre-synaptic terminal loss, and dendritic spines loss, lead to the primary stage of AD, introducing memory dysfunction [24,25,26].
Both neurodegenerative diseases, delirium, and AD, have a complex interrelationship among their pathophysiological mechanisms in which they can act interactively, independently, and simultaneously [27]. Delirium has been treated as a vulnerability marker for AD which can alter the potential neuronal injury that leads to AD. The risk of incident dementia is considerably increased by delirium, frequently misdiagnosed or confused with AD. Studies suggest that 22%-89% of patients with dementia experience delirium during critical medical events [28]. The outcomes of studies examining delirium-related biomarkers in people with AD have been mixed. However, there are associations between delirium and AD-associated biomarkers, suggesting that the underlying AD pathology might impact the development of delirium [29, 30]. Moreover, AD patients with delirium are at greater risk of suffering negative consequences, including death or being admitted to a nursing home and experiencing hastened cognitive loss [31]. Glucose utilization and insulin signaling are significantly decreased in AD patients [32, 33], which are also linked with delirium [34, 35].
This introduction indicates a complex and mysterious relationship between the etiology of delirium and AD. Even though a few weakly powered genetic investigations have been carried out, no persistent potential genes linked to delirium risk have been found [35,36,37]. To better understand the underlying pathophysiology of both diseases, we consider that conducting an in-depth investigation is necessary to decipher the common biomarkers and signaling pathways shared by both diseases. In addition, the conjugial nature of these two neurological diseases increases the challenge of developing successful treatments for both. Although very few drugs are being prescribed or treating the symptoms of AD [21], no specific drugs are being considered for delirium treatment, despite ongoing research [38,39,40,41].
We have carried out an integrated bioinformatics and system biology analysis to explore the typical potential molecular relationship between delirium and AD, exploring the common molecular signatures and pathways and the repurposable drug investigation for delirium and AD. The study was designed to capture the delirium pathophysiology associated with AD and dementia. The study also sought to elucidate the typical pathophysiological association between delirium and AD/dementia. The outcomes of this study were hypothesized to generate evidence for more profound knowledge and understanding of delirium and AD as well as better therapeutic development, especially for delirium.
Materials and methods
We used delirium-associated proteins and AD-related proteins from different independent sources, to identify the common proteomic biomarker candidates. Then, a network-based analysis approach was used to decipher the pathophysiological processes and their regulators. The entire study diagram has been presented in Fig. 1.
Data sources and descriptions
Due to a lack of delirium-gene expression data, the current study searched for delirium-associated gene expression data. For this study, the delirium-associated protein dataset was collected through a systematic literature review (SLR) (please see the Supplementary File 1 for details about the SLR) and the Comparative Toxicogenomics Database (CTD, http://ctdbase.org/) [42], a widely used database for investigating chemical genes or proteins relationships. In our study, we searched the proteins from CTD against delirium. Combining these two protein datasets, we have compiled a delirium-associated total protein seed dataset containing 524 unique gene encoded proteins (Table 1).
The AD-associated gene expression transcriptomic dataset was downloaded from the National Center for Biotechnology Information Gene Expression Omnibus (NCBI-GEO) data repository. The AD-associated microarray dataset (GSE36980 [45]) contained 80 samples of human postmortem brains, of which 33 samples were from AD-affected brains, and others were from non-AD brains collected from the area of the frontal cortex, temporal cortex, and hippocampus of brains.
Identification of differentially expressed genes (DEGs) between AD and control samples
We used the well-established linear model method for microarray data (LIMMA) [46] to identify AD-associated DEGs. The conventional LIMMA procedure employs an empirical Bayes estimation procedure to ‘moderate’ the ordinary t-test statistic by adjusting the sample variance using the distribution of all standard deviations. The p-values were adjusted using the Benjamini and Hochberg approach to control the multiple testing false discovery rate (FDR) [47]. Statistical significance was considered by an adjusted P-value < 0.05 and the |log2(FC)|> 0.5 (where FC means average fold change value) to identify the significant DEGs. The statistical tests were performed and implemented by the NCBI-GEO2R web tool (https://www.ncbi.nlm.nih.gov/geo/geo2r/, accessed on 22 June 2023).
In this statistical test, the kth gene (where k = 1,2, …, 27,925) was considered as DEG altering between the AD and non-AD groups if the adjusted Pk-value < 0.05 with the |log2(aFCk)|> 0.5 after controlling the FDR at 5% level or else, it was an equally expressed gene (EEG). If the kth gene's adjusted Pk-value were less than 0.05 and log2(aFCk) > 0.5 and log2(aFCk) < − 0.5, the gene was classified as either up- or down-regulated DEG, respectively. Here, the aFCk represents the average normalized fold change value of the kth gene’s expression arrays concerning AD and non-AD samples which can be defined as aFCk =\({\overline{x} }_{k}/{\overline{y} }_{k}\) (where \({\overline{x} }_{k} and\) \({\overline{y} }_{k}\) is the average value of the normalized expression array count of the kth gene in AD and non-AD samples). For instance, if the \({\overline{x} }_{k}\)=8 in the AD sample and \({\overline{y} }_{k}=2\), then the aFCk = 4, indicating that the kth gene is fourfold upregulated in AD compared to non-AD conditions and vice versa for downregulated genes.
Identification of shared DEGs (sDEGs) between AD and delirium
First, we collected DEGs between delirium and control samples from the literature. Suppose we denoted the delirium-associated DEGs encoded protein dataset by PD and the AD-associated DEGs encoded protein dataset by QAD. Then we selected the shared DEGs (sDEGs) encoded protein dataset between delirium and AD by Z = \({(P}_{D}\cap {Q}_{AD}\)) which has been used as the final combined analytical dataset for this study.
sDEGs-set enrichment (GSE) and annotation analysis
Gene ontology enrichment and functional signaling pathway analysis were conducted to identify the significant biological and molecular functions. The gene set enrichment and ontology analysis were performed using g:GOSt embedded in g:Profiler web server. The significant signaling pathways were retrieved from four databases including BioCarta, WikiPathways, KEGG, and Reactome. The significant signaling pathways and ontology terms were considered based on the adjusted P-value < 0.05 and the Benjamini and Hochberg [47] procedure for controlling FDR.
Identification of common key genes associated with delirium and AD
It is a common practice to investigate key proteins using protein–protein interaction (PPI) network analysis [48, 49]. In the current study, the STRING, a protein interactome database [50] was used to build the PPI network of the sDEGs encoded proteins between delirium and AD. The highly representative key common proteins, also known as hub-proteins, were retrieved using a topological investigation based on dual-metric measurement degree and betweenness on the PPI network using the Cytoscape [51]. ClueGO, a plug-in Cytoscape was used to create a network of functions for GO enrichment analysis utilizing the key hub genes under the statistical significance of P-value < 0.05.
Pre- and post-transcriptional gene regulatory network analysis
In this study, we have identified the pre- and post-transcriptional gene regulatory factors-microRNA (miRNA) and the transcriptional factor (TF) analyzing the interaction networks among the shared hub-genes encoded proteins and miRNAs and TFs, respectively. The interaction network of TFs and shared hub-DEGs was constructed using the JASPAR [52] TF database as well as the TarBase V8.0 and miRTarBase [53, 54] miRNAs databases were utilized to construct the miRNA-hub-DEGs interaction network. The Network Analyst [55] online server revealed all the regulatory networks. Dual-metric measurement degree and betweenness were implemented in network analysis and visualization.
Disease-gene interaction analysis
The disease multimorbidity association of the sDEGs was investigated using the disease-gene interaction network. To discover this, the DisGeNET database [56, 57] was utilized and then Cytoscape was used to analyze the network under the dual-metric topological measurement condition degree and betweenness to capture the important and significant disease interactions with the hub-DEGs. The significant diseases were also highlighted which are highly comorbid with delirium and AD.
Drug repurposing and molecular docking
Since we observed the internal pathophysiological relationship, we attempted to identify the potential repurposable drugs against delirium and AD using computational molecular docking analysis. The FDA-approved drugs that are used for neurological treatment (like AD) have been retrieved from the online drug-repositioning tool and database Connectivity Map (CMap) [58] (Supplementary File 2). The drug repurposing database was used to search the drug molecules against hub-proteins. Only the top-ranked associated repurposable drug molecules were collected for the docking analysis with our hub-proteins. Then we performed molecular docking analysis between the top-ranked repurposable drug molecules and target proteins as in-silico validation. The binding affinity of the repurposable drugs with the top key common hub-proteins and TFs was investigated by a molecular docking simulation study. The 3D structures of the drug target key hub-proteins, TF proteins, and repurposable drugs were downloaded from the Protein Data Bank (PDB) [59] and SWISS-MODEL [60], a homology modeling-based database as well as from PubChem database [61]. The 3D drug target proteins were visualized and preprocessed by removing the co-crystal ligands and water molecules using Discovery Studio Visualizer 2019 [55], Swiss PDB Viewer Software, and MGLTools Software [62]. The energy minimization of the drug compounds was performed by applying the MMFF94 force field [63]. Molecular docking analysis was performed using AutoDockTools 4.2 [64] and AutoDock Vina [65]. For the drug-protein interaction, the highest docking score with the best-fit posture was considered to select the best repurposable drug for delirium as well as AD.
Molecular Dynamic (MD) simulations
The MD simulation was employed to explore the dynamic nature of the top-ranked protein-drug complex by using the YASARA Dynamics software [66], and the AMBER14 force field [67]. The simulation included the top four protein-drug complexes, ALB-Donepezil, ALB-Tacrine, GAD2-Donepezil, and GAD2-Tacrine from molecular docking analysis. The hydrogen bonding network of the target protein-drug complex was optimized and solvated using a TIP3P [68] water model in a simulation cell before simulation. Given a solvent density of 0.9971 gL-1, periodic boundary conditions were maintained. The ALB-Donepezil, ALB-Tacrine, GAD2-Donepezil, and GAD2-Tacrine complex contained 115,942, 114,258, 158,429, and 159,893 atoms respectively at the initial energy minimization process using the simulated annealing method with the steepest gradient approach (5000 cycles). Under physiological circumstances (298 K, pH 7.4, 0.9% NaCl) [69], every simulation was run with a repeated time-step method [70] utilizing a time-step interval of 2.50 fs. The linear constraint solver (LINCS) [71] algorithm was used to limit all bond lengths and SETTLE [72] was used for water molecules. The root-mean-squared deviation (RMSD) and molecular mechanics Poisson-Boltzmann surface area (MM-PBSA) binding free energy was calculated for up to 100 ns MD simulation under the Berendsen thermostat [73] and constant pressure. The analysis was performed using the default script of YASARA macro and SciDAVis software available at http://scidavis.sourceforge.net/. The MM-PBSA binding free energy was calculated by the following equation [74], using YASARA built-in macros using AMBER 14 as a force field, with larger positive energies indicating better binding [75],
Results
Identification of DEGs between AD and control samples
The statistical analysis revealed a total of 2257 DEGs (i.e., 534 up-regulated and 1723 down-regulated genes) with their official gene symbol identified from the gene expression data analysis. The significant upregulated and downregulated genes and their mean expression difference (AD vs. non-AD) plot are shown in Fig. 2A and B respectively.
Identification of DEGs between delirium and control samples
A total of 189 unique delirium-associated genes and their encoded proteins were found from 78 included studies (Table 1) in the comprehensive SLR. Under the cutoff inference score (> 40), the CTD database revealed a total of 350 delirium-associated gene encoded proteins. Then, the two protein datasets from separate sources (SLR and CDT database) were combined (mathematical union) to create an integrated delirium-associated protein dataset with 524 unique proteins (Table 1).
Identification of sDEGs between AD and delirium
A total of 99 common shared DEGs (sDEGs) between delirium and AD were identified as displayed in Fig. 2C. The distribution of up and downregulated genes with the delirium-associated genes shows that a total of 79 downregulated genes and 20 upregulated genes were common between delirium and AD (Table 2). The shared genes and their encoded proteins were utilized to identify the common regulatory biomolecules and common pathophysiological relationships between delirium and AD in the downstream analysis.
sDEGs set enrichment analysis with GO-terms and pathways
Based on the statistical significance criteria (AdjP-value < 0.05) under the controlled FDR, the top significant functional pathways and enriched GO shared by the common genes of delirium and AD are shown in Fig. 3. The bubble plots in Fig. 3A and B have been constructed from g:GOSt server, representing the top significant GO terms (Fig. 3A) and the functional pathways (Fig. 3B). The analysis revealed the significant GO terms including the biological process (BP) molecular functions (MF) and cellular components (CC). The most significant GO terms are represented in Fig. 3A. According to the GO enrichment analysis, the chemical synaptic transmission, cell communication, memory, anterograde trans-synaptic signaling, response to stimulus, protein-binding activity, neurotransmitter receptor activities, and other synoptic signaling activities are highly enriched and are significant GO terms commonly shared by delirium and AD-associated genes (Fig. 3A).
The entire neurotransmission system along with the different signaling pathways were highly enriched pathways among the common genes of delirium and AD. The enriched functional pathways commonly linked with delirium and AD that have been identified, are mostly associated with the nervous system and their signaling synapse mechanism as well as the chemical reaction of receptors with ligand chemical molecules (Fig. 3). For example, among the most significant enriched pathways shared by delirium and AD, the transmission across chemical synapses, pathways of neurodegeneration-multiple diseases, signal transduction, neuroinflammation, and glutamatergic signaling pathway, brain-derived neurotrophic factor (BDNF) signaling pathway, fragile-X syndrome, oxidative stress, and hypoxia associated pathways are the most important. The detailed analysis output of GSE and functional pathway analysis has been provided in Supplementary File 3.
PPI network analysis of sDEGs
The PPI network of the sDEGs-encoded proteins revealed the most highly connected shared key proteins which are also known as drug-targeted hub-proteins. Among the signature proteins, three key proteins (ALB, AGT and GFAP) were AD-associated upregulated, and the others were found downregulated. The top 10 hub-proteins including ALB, APP, BDNF, CREB1, DLG4, GAD1, GAD2, GFAP, GRIN2B and GRIN2A were found in the PPI network (Fig. 4) and utilized for further downstream analysis.
Figure 4 demonstrated that the two upregulated and five downregulated gene-encoded proteins were not interconnected with any other proteins in the network (Fig. 4). With the PPI network, we observed the engagement and association of the signature proteins with delirium and AD. The ClueGO-derived GO network also revealed the association of mental dysfunction-related pathways and biological function which were consistent with the overall GSE analysis (Supplementary Fig. 1).
The sDEGs regulatory network analysis
The common sDEGs-TFs and DEGs-miRNAs interaction network is presented in Fig. 5. The gene-TFs regulatory network showed the key regulatory TF including the FOXC1, GATA2, and FOXL1. The highly connected TFs are represented in Fig. 5A with the green diamond nodes. The common DEGs-miRNA network analysis revealed the potential key miRNAs namely, miR-16-5p, miR-1-3p, and miR-34a-5p (Fig. 5B).
Disease-gene network fetched the key neurological disorders associated with the shared DEGs
The interaction network analysis revealed the associated diseases with the common hub-genes associated with delirium and AD. The interaction network is displayed in Fig. 6. Most importantly Alzheimer’s disease 2, dementia, cognitive disorders, Parkinson's disease and disorders, and brain diseases were the most significant diseases associated with common hub-genes (Fig. 6). The disease interaction and the association indicate that neurological complications and disorders are highly comorbid with delirium and AD development.
The hub-proteins guided drug discovery
In this part the top 10 common hub-DEGs encoded hub-proteins and 3 key TFs (total = 13) proteins for molecular docking simulation were considered. The 3D-structure of the 10 hub-proteins including ALB, BDNF, GRIN2B, CREB1, APP, DLG4, GFAP, GAD1, GRIN2A and GAD2 were collected from the PDB database using the codes 7VR0, 1BND, 7EU8, 5ZKO, 1AAP, 6SPV, 6A9P, 3VP6, 5H8Q, and 2OKK respectively. The 3D structure of GATA2 TF-protein was also downloaded from the PDB database using the code 5O9B whereas the other two 3D structures of TFs FOXC1 and FOXL1 were collected from the SWISS-MODEL using UniProt with IDs Q12948 and Q12952. The 3D structure of eight FDA-approved neurological drugs was collected from the PubChem database and used for molecular docking against the drug receptor proteins associated with delirium and AD. Based on the binding affinity scores (kcal/mol) between the receptor proteins and the drug agents, the top repurposable drug molecules and the most effective drug targets were confirmed. The affinity scores were ordered and plotted in a heatmap against the receptor proteins in Fig. 7. In our investigation, the ALB and GAD2 were found as leading and prominent drug target receptor proteins associated with delirium and AD where donepezil (with ALB: -8.8 kcal/mol and with GAD2: -9.0 kcal/mol) and tacrine (with ALB: -8.0 kcal/mol) showed the maximum binding affinity scores with the two target proteins compared to others lead components (Fig. 7). The docking analysis revealed that most of the drug agents performed well with the target proteins which resulted in the GRIN2B protein also docking well with citicoline (-7.5 kcal/mol) and tacrine (-7.6 kcal/mol) drug molecules. The details docking score matrix is provided in Supplementary File 4.
Table 3 represents the gist of the molecular docking interaction summary of our top drug target proteins (ALB and GAD2) with the prominent drug candidates (Tacrine and Donepezil) scoring maximum binding affinity. The best docking pose (3D) of the drug molecule, the interaction complex (2D), and the adjacent interacting residues along with the bond and distance (Å) are reported in Table 3. The interaction pose of the target receptor and drug molecule indicates that the drug molecule fits on the target protein’s pocket with significant binding affinities.
MD simulation
The complex stability analysis through MD simulation between the top-ranked drug target and drug molecules showed significant stability between the initial drug target and complex moving variation over the 100 ns MD-PBSA simulation. Fig. 8A shows the calculated RMSD for all four protein-drug complexes ALB-Donepezil, ALB-Tacrine, GAD2-Donepezil, and GAD2-Tacrine.
The system provided an average RMSD of 2.185 Å (Range: 0.428 Å to 3.148 Å), 2.255 Å (Range: 0.449 Å to 3.177 Å), 7.540 Å (Range: 0.483 Å to 8.434 Å) and 7.590 Å (Range: 0.504 Å to 8.727 Å) for the ALB-Donepezil, ALB-Tacrine, GAD2-Donepezil and GAD2-Tacrine respectively. The GAD2 complex structures fluctuated for the drug molecules up to 15 ns and became stable during the remaining simulation. The RMSD plot indicates that the ALB complexes were more stable with the reported drugs during the entire simulation than GAD2. The MM-PBSA binding energy for four complexes shows the average binding energy 307.061 kJ/mol, 192.694 kJ/mol, 105.350 kJ/mol, and 111.743 kJ/mol for ALB-Donepezil, ALB-Tacrine, GAD2-Donepezil, and GAD2-Tacrine complex respectively (Fig. 8B).
Discussion
This study has focused on deciphering the interactions and pathophysiological pathways shared by the shared key proteomic biomarkers between delirium and AD. The significant AD-associated DEGs were compared with the delirium-associated genes to identify the common genomic signatures and found 99 common genes between the two conditions. Among the common genes shared by delirium and AD, it was observed that the common genes are differentially expressed in AD where 20 genes were upregulated and 79 were downregulated.
The shared functional pathways between the two diseases show the epidemiological and internal pathophysiological relationship between them. For instance, transmission across chemical synapses is one of the most important and significant pathways shared by the common genes for transferring chemical neurotransmitters across the neurons [76]. One neuron can quickly and efficiently stimulate or inhibit the neuronal activity of another neuron via chemical synapses. The neurodegeneration pathways caused by multiple diseases play a significant role in enhancing the progression of memory loss and ultimately developing distinct brain-dysfunctional diseases like AD, dementia, and Parkinson’s disease [77]. As the common genes of delirium and AD were significantly enriched in this pathway, it indicates that delirium is also associated with permanent cognitive and/or motor dysfunction, supported by different studies [27, 78]. Another significant shared pathway was the neuroinflammation and glutamatergic signaling pathway. Neuroinflammation is considered the primary process that triggers delirium under any critical medical condition. The neuronal and synaptic dysfunction is changed due to the neuroinflammatory, which is after the abnormal neurobehavioral and mental disorder symptoms [79, 80]. The common genes also enriched the BDNF signaling pathway, which plays a vital role in the pathogenesis of neurodegenerative diseases by accelerating TrkB-mediated neuronal events [81]. Oxidative stress is a significant mechanism linked to chronic inflammation and age-related disorders that may also be connected to delirium pathogenesis [82]. Among the other important pathways, the glutamatergic signaling in the central nervous system, and cholinergic functional pathways are crucial pathways associated with delirium and AD development [83]. Studies show that anesthesia drugs during surgery can directly act on the central cholinergic system which leads to postoperative delirium and mental dysfunction [84]. Drugs used for anesthesia could be one of the major factors for postoperative delirium [85, 86] and it demands rigorous research to decipher the molecular interaction of the drug molecules with the delirium-associated target proteins. The pathway indicates the interrelationship of neuronal activity, cognitive impairment, and conditions like AD and dementia. Investigating these pathways enhances the influence of delirium on chronic brain diseases.
The PPI network analysis of sDEGs-encoded proteins revealed the top key hub-proteins where most of them came from the down-regulated genes. Among the top ten hub-proteins, the ALB and GFAP were upregulated from an AD perspective and others were downregulated. The expression profiles of the common genes between delirium and AD indicate that delirium superimposed on AD might also be triggered by the downregulated genes associated with delirium. Among the key hub-genes, albumin (ALB) is considered a potential biomarker to diagnose delirium among surgical patients [87]. The studies support that lower albumin levels are highly associated with delirium development postoperatively [88,89,90,91]. The astrocyte-expressed glial fibrillary acidic protein (GFAP) [92], is used to identify astrocytosis in cases of neurodegeneration and is associated with traumatic brain injury that might involve mental illness like AD. Among the other key downregulated hub-genes, BDNF, GRIN2B, and CREB1 are highly connected with the other genes. The increased protein level of BDNF is associated with delirium diagnosis and quick recovery from postoperative delirium [93,94,95] whereas it has been found in lower levels among AD patients [96]. Studies suggested that genetic variation of the GRIN2B gene might be associated with the molecular mechanism of AD [97, 98] and its molecular variations may offer a crucial tip for understanding the molecular causes of AD [99]. The GSE analysis of the common genes revealed the cMAP-signaling pathway, mainly enhanced by CREB1 genes associated with mental depression [100]. The pathway enrichment analysis of the hub-genes using ClueGo revealed the important pathways, namely BDNF-TrkB signaling pathway, NMDA glutamate receptor activity, synaptic signaling pathways, fragile X syndrome, cocaine addiction, and amphetamine addiction (Supplementary Fig. 1). The pathways are aligned with the overall GSE analysis results. The top significant hub-genes play a substantial role in delirium development and AD. The genes might influence higher delirium occurrence with the concurrence of critical medical conditions of AD-affected patients. Therefore, the hub-genes can serve as a potential biomarker to diagnose delirium and could be treated as a potential therapeutic target for drug development.
The gene regulatory network (GRN) analysis detected some key transcription factors (TFs), namely FOXC1, GATA2, and FOXL1, as the transcriptional regulators of shared key genes as well as key miRNAs, notably miR-16-5p, miR-1-3p, and miR-34a-5p as the post-transcriptional regulators. Neuroinflammation and neuronal death are linked to the FOXC1 TF whereas, neurodegenerative consequences including Alzheimer's disease, dementia, and Parkinson's disease are strongly correlated with neuroinflammation [101,102,103]. The key TF GATA2 is associated with the Neuroglobin (NGB) gene expression when the neural disease (like AD) is connected with the expression level of the NGB gene [104]. The results suggest that FOXL1, a transcriptional repressor regulates the development of the central nervous system in Zebra fish [105] and is also associated with AD [106] as reported in previous studies. The miR-16-5p miRNA plays a significant role in neuronal cell apoptosis in AD [107, 108]. The miR-1-3p is directly involved in Fas Apoptotic Inhibitory Molecule (FAIM) expression which is closely associated with the physiological and pathological processes of Alzheimer’s and Parkinson’s diseases [109]. Different clinical and molecular studies suggested that the plasma level of miR-34a-5p miRNA is being considered as an early biomarker [110, 111], which also decreases oxidative stress and apoptosis condition by inhibiting β-amyloid (Aβ)-induced neurotoxicity in AD [112]. The EGR1 TF regulates the AChE expression contributing to cholinergic function alteration in AD development [113] which may contribute to delirium as well. The SP1 is known as a pro-inflammatory TF that regulates the AD causal genes including amyloid precursor protein (APP) and β-secretase (BACE1) gene expression [114, 115]. Besides the TF, the miR-103 and miR-107 are directly associated with neurodegenerative diseases like AD and neurodegeneration-associated pathways which play an important role in delirium as well [116, 117]. Since the regulatory molecules are associated with the common hub-genes of delirium and AD their functionalities are directly involved with neurological complications and pathophysiology and they might have close connectivity with delirium occurrence and development.
The disease-gene interaction network revealed the comorbidities associated with delirium and AD. The common genes associated with diseases contain mental dysfunctions and disorder-related complications that might influence delirium and AD development, or they could boost medical complications in patients with delirium and AD. The comorbidity analysis revealed alcoholic intoxication which could lead to delirium. The results are consistent with the GSE and ClueGO GO group analysis results including cocaine and nicotine addiction pathways. Studies suggested that the alcohol withdrawal/alteration could result in delirium [118,119,120] where the key hub-genes may have been involved.
Repurposable drugs are considered a great source of treatment in any emergency. In this aspect, the FDA-approved neurological drugs especially used for AD treatment were retrieved from the CMap database. The computational molecular docking simulation study was implemented to investigate the drug target properties of our proposed drug target proteins which might be investigated further for more effective therapeutic development against delirium and AD. The docking analysis revealed that the drug molecules significantly interact with the target protein pockets. Among the drug target proteins, ALB and GAD2 were found to be highly interacting drug targets compared to others. Tacrine and donepezil showed the highest binding affinity scores which indicate the primary properties of drug candidate molecules. Tacrine was approved by the FDA as one of the first drugs to treat AD [121] although it has been prescribed with limitations for easing AD symptoms [122, 123]. Studies suggest that tacrine-related drugs could be a potential source for AD treatment [124]. Tacrine also showed effective improvement in treatment of cholinergic delirium [125]. On the other hand, donepezil, an acetylcholinesterase (AChE) inhibitor has already been investigated as a drug for neurological or psychiatric complications including delirium and AD [126, 127]. Studies supported that the donepezil showed strong significant binding affinity with AChE which triggers the cholinergic pathways on AD therapeutics [128,129,130]. Clinical improvement was investigated by using the donepezil for the Alzheimer dementia patients [131]. Research indicates that donepezil has being investigated as a potential medication for treating delirium [132, 133]. Donepezil medication also improved the critical condition of dementia patients and reduced the delirium development [134]. The computational analysis in this study revealed consistent findings about donepezil as a prominent therapeutic candidate which will influence the therapeutic development for delirium as well. The 100 ns MD-based simulation revealed the stability of the reported protein-drug complexes suggesting significant structural consistency according to the physical law [135]. Based on the prominent properties of our proposed drug molecules, further clinical and pharmacological research is needed for effective therapeutic development against delirium and AD targeting the key drug target biomolecules reported in this study.
Strength and limitations
The current study utilized AD-associated gene expression data collected from human brain samples, which could explain a greater genomic signature than blood and other tissue samples. This study collected a comprehensive delirium-associated gene encoded protein dataset which included an SLR and an independent CTD database that can be explored for further delirium research. The study outlined key drug-target biomolecules including hub-proteins, TFs, and miRNAs which are jointly functional in delirium and AD. These important biomarker genes will open a new dimension of research in diagnosis, prognosis, and therapeutic development. The proposed repurposable neurological drugs showed significant binding affinity against the therapeutic targets which augers well for therapeutic development for delirium and AD.
While the study identified common molecular signatures, it may not fully elucidate the intricate biological pathways linking delirium and AD. Further research is needed to gain a comprehensive understanding of the shared pathophysiological processes. There might be some inconsistencies about the common molecular functionality between delirium and AD, since AD is highly accountable for dementia whereas delirium is generally a short-term cognitive impairment. Therefore, the common pathophysiological functions between delirium and AD would be considered when they act conjugately. In this aspect, the identified key proteins and their functionalities may differ for independent delirium episodes and their subtypes. Predominantly, both neurological conditions are highly prevalent among older patients and have a great chance to be comorbid to each other when occur together. Although the present study reported several important key proteins, further research needs to be conducted to identify a single biomarker of delirium and AD. The study's suggestions for drug repurposing for delirium might be influenced by bias or limited data availability. Rigorous clinical trials are needed to validate the effectiveness and safety of repurposed drugs. Moreover, the delirium associated gene expression data should be generated to elucidate the genetic engagement on disease pathophysiology.
Implementation
This study aimed to understand the pathophysiological relationship between delirium and AD along with identifying potential drug targets and repurposable drug candidates. The outcomes will significantly contribute to a better understanding of the common key genomic biomarkers and shared signaling pathways associated with delirium and AD. This will enrich the pathophysiological knowledge about delirium and AD, their cooccurrence, and also delirium superimposed on dementia. Healthcare policies should prioritize biomarker-based early detection of individuals at risk for both delirium and AD. The comorbidity analysis associated with delirium and AD reported significant symptoms which will contribute to healthcare practitioner knowledge for good practice of diagnosis, monitoring, and management of delirium and AD. Healthcare systems might adopt integrated care models that bring together specialists in geriatric medicine, neurology, psychiatry, and genetics to comprehensively address the overlapping risk factors and underlying genetic connections between delirium and AD. Finally, if the reported repurposable drugs are considered for in-depth clinical investigation for further validation, they will be a potential source for enhancing precision medicine and the process of therapeutic development against delirium and AD. This information could guide personalized risk assessment and early intervention strategies.
Conclusion
The literature supports that delirium individually is a common phenomenon among older patients and significantly increases the economic burden, mortality, and morbidity. When AD-affected patients develop delirium in a medical setting, the consequences are more severe. This study identified several significant biomarker proteins such as ALB, BDNF, GRIN2A, and GAD2 as potential candidates for diagnosis, prognosis, and therapeutic development against delirium and AD. The transmission across chemical synapses, neurodegeneration, signal transduction, neuroinflammation and glutamatergic signaling pathway, BDNF signaling pathway, fragile-X syndrome, oxidative stress, hypoxia, and cholinergic functional pathways were most significantly associated with delirium and AD-associated pathophysiology. Moreover, the MD analysis and simulation study (100 ns) among the common hub-proteins and neurological repurposable drugs provided the top-ranked drug candidates (tacrine and donepezil) and the prominent drug target proteins (ALB and GAD2), significant for therapeutic development against delirium and AD. The findings were consistent with and supported by the outcomes of previous studies as we discussed earlier. The findings of this study will strengthen the molecular research foundation for delirium and AD pathophysiological mechanisms. Furthermore, the reported drug targets and the drug molecules will enhance efficient therapeutic development for delirium and AD by further validation under in-depth pharmacological and clinical research.
Availability of data and materials
All the data used in this study is publicly available secondary data.
References
Williams ST, Dhesi JK, Partridge JSL. Distress in delirium: causes, assessment and management. Eur Geriatr Med. 2020;11:63–70. https://doi.org/10.1007/s41999-019-00276-z.
Dunne SS, Coffey JC, Konje S, Gasior S, Clancy CC, Gulati G, et al. Biomarkers in delirium: a systematic review. J Psychosom Res. 2021;147:110530. https://doi.org/10.1016/j.jpsychores.2021.110530.
AGS/NIA Delirium Conference Writing Group PC and F. The American geriatrics society/national institute on aging bedside-to-bench conference: research agenda on delirium in older adults. J Am Geriatr Soc. 2015;63:843–52. https://doi.org/10.1111/jgs.13406.
Marcantonio ER. Delirium in hospitalized older adults. Solomon CG, editor. N Engl J Med. 2017;377:1456–66. https://doi.org/10.1056/NEJMcp1605501.
Poulsen LM, Estrup S, Mortensen CB, Andersen-Ranberg NC. Delirium in Intensive Care. Curr Anesthesiol Rep. 2021;11:516–23. https://doi.org/10.1007/s40140-021-00476-z.
Sepulveda E, Franco JG, Trzepacz PT, Gaviria AM, Meagher DJ, Palma J, et al. Delirium diagnosis defined by cluster analysis of symptoms versus diagnosis by DSM and ICD criteria: Diagnostic accuracy study. BMC Psychiatry. 2016;16:167. https://doi.org/10.1186/s12888-016-0878-6.
Persico I, Cesari M, M, Morandi A, Haas J, Mazzola P, Zambon A, et al. Frailty and Delirium in Older Adults: A Systematic Review and Meta-Analysis of the Literature. J Am Geriatr Soc. 2018;66(10):2022–30. https://doi.org/10.1111/jgs.15503.
Gross AL, Jones RN, Habtemariam DA, Fong TG, Tommet D, Quach L, et al. Delirium and Long-term Cognitive Trajectory Among Persons With Dementia. Arch Intern Med. 2012;172:1324. https://doi.org/10.1001/archinternmed.2012.3203.
Silver G, Traube C, Gerber LM, Sun X, Kearney J, Patel A, et al. Pediatric delirium and associated risk factors: a single-center prospective observational study. Pediatr Crit Care Med. 2015;16:303–9. https://doi.org/10.1097/PCC.0000000000000356.
Smith PJ, Attix DK, Weldon BC, Greene NH, Monk TG. Executive function and depression as independent risk factors for postoperative delirium. Anesthesiology. 2009;110:781–7. https://doi.org/10.1097/ALN.0b013e31819b5bc2.
Wilson K, Broadhurst C, Diver M, Jackson M, Mottram P. Plasma insulin growth factor - 1 and incident delirium in older people. Int J Geriatr Psychiatry. 2005;20:154–9. https://doi.org/10.1002/gps.1265.
Velayati A, Vahdat Shariatpanahi M, Shahbazi E, Vahdat SZ. Association between preoperative nutritional status and postoperative delirium in individuals with coronary artery bypass graft surgery: a prospective cohort study. Nutrition. 2019;66:227–32. https://doi.org/10.1016/j.nut.2019.06.006.
Sanford AM, Flaherty JH. Do nutrients play a role in delirium? Curr Opin Clin Nutr Metab Care. 2014:45–50. https://doi.org/10.1097/MCO.0000000000000022.
Inouye SK, Westendorp RGJ, Saczynski JS. Delirium in elderly people. Lancet. 2014:911–922. https://doi.org/10.1016/S0140-6736(13)60688-1.
Taccone FS, Su F, Pierrakos C, He X, James S, Dewitte O, et al. Cerebral microcirculation is impaired during sepsis: an experimental study. Crit Care. 2010;14:R140. https://doi.org/10.1186/cc9205.
Bendahan N, Neal O, Ross-White A, Muscedere J, Boyd JG. Relationship Between Near-Infrared Spectroscopy-Derived Cerebral Oxygenation and Delirium in Critically Ill Patients: A Systematic Review. J Intensive Care Med. 2019;34:514–20. https://doi.org/10.1177/0885066618807399.
van den Boogaard M, Kox M, Quinn KL, van Achterberg T, van der Hoeven JG, Schoonhoven L, et al. Biomarkers associated with delirium in critically ill patients and their relation with long-term subjective cognitive dysfunction; indications for different pathways governing delirium in inflamed and noninflamed patients. Crit Care. 2011;15:R297. https://doi.org/10.1186/cc10598.
Vasunilashorn SM, Dillon ST, Inouye SK, Ngo LH, Fong TG, Jones RN, et al. High C-reactive protein predicts delirium incidence, duration, and feature severity after major noncardiac surgery. J Am Geriatr Soc. 2017;65:e109–16. https://doi.org/10.1111/jgs.14913.
Hall RJ, Watne LO, Cunningham E, Zetterberg H, Shenkin SD, Wyller TB, et al. CSF biomarkers in delirium: a systematic review. Int J Geriatr Psychiatry. 2018;33:1479–500. https://doi.org/10.1002/gps.4720.
Ali S, Patel M, Jabeen S, Bailey RK, Patel T, Shahid M, et al. Insight into delirium. Innov Clin Neurosci. 2011:25–34. Available: http://www.ncbi.nlm.nih.gov/pubmed/22132368.
Breijyeh Z, Karaman R. Comprehensive review on Alzheimer’s disease: causes and treatment. Molecules. 2020:5789. https://doi.org/10.3390/MOLECULES25245789.
Spires-Jones TL, Hyman BT. The intersection of amyloid beta and tau at synapses in Alzheimer’s disease. Neuron. 2014:756–771. https://doi.org/10.1016/j.neuron.2014.05.004.
Singh SK, Srivastav S, Yadav AK, Srikrishna S, Perry G. Overview of Alzheimer’s disease and some therapeutic approaches targeting A β by using several synthetic and herbal compounds. Oxid Med Cell Longev. 2016;2016:1–22. https://doi.org/10.1155/2016/7361613.
Overk CR, Masliah E. Pathogenesis of synaptic degeneration in Alzheimer’s disease and Lewy body disease. Biochem Pharmacol. 2014:508–516. https://doi.org/10.1016/j.bcp.2014.01.015.
Lleó A, Núñez-Llaves R, Alcolea D, Chiva C, Balateu-Paños D, Colom-Cadena M, et al. Changes in synaptic proteins precede neurodegeneration markers in preclinical Alzheimer’s disease cerebrospinal fluid*. Mol Cell Proteomics. 2019;18:546–60. https://doi.org/10.1074/mcp.RA118.001290.
Tarawneh R, D’Angelo G, Crimmins D, Herries E, Griest T, Fagan AM, et al. Diagnostic and prognostic utility of the synaptic marker neurogranin in Alzheimer disease. JAMA Neurol. 2016;73:561–71. https://doi.org/10.1001/jamaneurol.2016.0086.
Fong TG, Vasunilashorn SM, Libermann T, Marcantonio ER, Inouye SK. Delirium and Alzheimer disease: a proposed model for shared pathophysiology. Int J Geriatr Psychiatry. NIH Public Access. 2019:781–789. https://doi.org/10.1002/gps.5088.
Fick DM, Agostini J V., Inouye SK. Delirium superimposed on dementia: a systematic review. J Am Geriatr Soc. 2002:1723–1732. https://doi.org/10.1046/j.1532-5415.2002.50468.x.
Fong TG, Inouye SK. The inter-relationship between delirium and dementia: the importance of delirium prevention. Nat Rev Neurol. 2022:579–596. https://doi.org/10.1038/s41582-022-00698-7.
Idland AV, Wyller TB, Stoen R, Eri LM, Frihagen F, Ræder J, et al. Preclinical amyloid-β and axonal degeneration pathology in delirium. Zerr I, editor. J Alzheimer’s Dis. 2016;55:371–9. https://doi.org/10.3233/JAD-160461.
Oh ES, Fong TG, Hshieh TT, Inouye SK. Delirium in older persons: advances in diagnosis and treatment. JAMA - J Am Med Assoc. 2017:1161–1174. https://doi.org/10.1001/jama.2017.12067.
Meltzer CC, Zubieta JK, Brandt J, Tune LE, Mayberg HS, Frost JJ. Regional hypometabolism in Alzheimer’s disease as measured by positron emission tomography after correction for effects of partial volume averaging. Neurology. 1996;47:454–61. https://doi.org/10.1212/WNL.47.2.454.
Hölscher C. Insulin signaling impairment in the brain as a risk factor in Alzheimer’s disease. Front Aging Neurosci. 2019. https://doi.org/10.3389/fnagi.2019.00088.
Kealy J, Murray C, Griffin EW, Lopez-Rodriguez AB, Healy D, Tortorelli LS, et al. Acute inflammation alters brain energy metabolism in mice and humans: role in suppressed spontaneous activity, impaired cognition, and delirium. J Neurosci. 2020;40:5681–96. https://doi.org/10.1523/JNEUROSCI.2876-19.2020.
Wilson JE, Mart MF, Cunningham C, Shehabi Y, Girard TD, MacLullich AMJ, et al. Delirium. Nat Rev Dis Prim. 2020;6:90. https://doi.org/10.1038/s41572-020-00223-4.
McCoy TH, Hart K, Pellegrini A, Perlis RH. Genome-wide association identifies a novel locus for delirium risk. Neurobiol Aging. 2018;68(160):e9-160.e14. https://doi.org/10.1016/j.neurobiolaging.2018.03.008.
Adamis D, Meagher D, Williams J, Mulligan O, McCarthy G. A systematic review and meta-analysis of the association between the apolipoprotein E genotype and delirium. Psychiatr Genet. 2016;26:53–9. https://doi.org/10.1097/YPG.0000000000000122.
Agar MR, Lawlor PG, Quinn S, Draper B, Caplan GA, Rowett D, et al. Efficacy of oral risperidone, haloperidol, or placebo for symptoms of delirium among patients in palliative care: a randomized clinical trial. JAMA Intern Med. 2017;177:34–42. https://doi.org/10.1001/jamainternmed.2016.7491.
Girard TD, Exline MC, Carson SS, Hough CL, Rock P, Gong MN, et al. Haloperidol and ziprasidone for treatment of delirium in critical illness. N Engl J Med. 2018;379:2506–16. https://doi.org/10.1056/nejmoa1808217.
Page VJ, Ely EW, Gates S, Zhao XB, Alce T, Shintani A, et al. Effect of intravenous haloperidol on the duration of delirium and coma in critically ill patients (Hope-ICU): a randomised, double-blind, placebo-controlled trial. Lancet Respir Med. 2013;1:515–23. https://doi.org/10.1016/S2213-2600(13)70166-8.
Van Den Boogaard M, Slooter AJC, Brüggemann RJM, Schoonhoven L, Beishuizen A, Vermeijden JW, et al. Effect of haloperidol on survival among critically ill adults with a high risk of delirium the REDUCE randomized clinical trial. JAMA - J Am Med Assoc. 2018;319:680–90. https://doi.org/10.1001/jama.2018.0160.
Davis AP, Wiegers TC, Johnson RJ, Sciaky D, Wiegers J, Mattingly CJ. Comparative Toxicogenomics Database (CTD): update 2023. Nucleic Acids Res. 2023;51:D1257–62. https://doi.org/10.1093/nar/gkac833.
Moher D, Liberati A, Tetzlaff J, Altman DG, Altman D, Antes G, et al. Preferred reporting items for systematic reviews and meta-analyses: the PRISMA statement. PLoS Med. 2009:e1000097. https://doi.org/10.1371/journal.pmed.1000097.
Takahashi Y, Terada T, Muto Y. Systems level analysis and identification of pathways and key genes associated with delirium. Genes (Basel). 2020;11:1225. https://doi.org/10.3390/genes11101225.
Hokama M, Oka S, Leon J, Ninomiya T, Honda H, Sasaki K, et al. Altered expression of diabetes-related genes in Alzheimer’s disease brains: the Hisayama study. Cereb Cortex. 2014;24:2476–88. https://doi.org/10.1093/cercor/bht101.
Gentleman RC, Carey VJ, Bates DM, Bolstad B, Dettling M, Dudoit S, et al. Bioconductor: open software development for computational biology and bioinformatics. Genome Biol. 2004;5:R80. https://doi.org/10.1186/gb-2004-5-10-r80.
Benjamini Y, Hochberg Y. Controlling the false discovery rate: a practical and powerful approach to multiple testing. J R Stat Soc Ser B. 1995;57:289–300. https://doi.org/10.1111/j.2517-6161.1995.tb02031.x.
Mosharaf MP, Kibria MK, Hossen MB, Islam MA, Reza MS, Mahumud RA, et al. Meta-data analysis to explore the hub of the hub-genes that influence SARS-CoV-2 infections highlighting their pathogenetic processes and drugs repurposing. Vaccines. 2022;10:1248. https://doi.org/10.3390/vaccines10081248.
Mosharaf MP, Reza MS, Gov E, Mahumud RA, Mollah MNH. Disclosing potential key genes, therapeutic targets and agents for non-small cell lung cancer: evidence from integrative bioinformatics analysis. Vaccines. 2022;10:771. https://doi.org/10.3390/vaccines10050771.
Szklarczyk D, Morris JH, Cook H, Kuhn M, Wyder S, Simonovic M, et al. The STRING database in 2017: quality-controlled protein-protein association networks, made broadly accessible. Nucleic Acids Res. 2017;45:D362–8. https://doi.org/10.1093/nar/gkw937.
Shannon P, Markiel A, Ozier O, Baliga NS, Wang JT, Ramage D, et al. Cytoscape: a software environment for integrated models of biomolecular interaction networks. Genome Res. 2003;13:2498–504. https://doi.org/10.1101/gr.1239303.
Khan A, Fornes O, Stigliani A, Gheorghe M, Castro-Mondragon JA, van der Lee R, et al. JASPAR 2018: update of the open-access database of transcription factor binding profiles and its web framework. Nucleic Acids Res. 2018;46:D260–6. https://doi.org/10.1093/nar/gkx1126.
Karagkouni D, Paraskevopoulou MD, Chatzopoulos S, Vlachos IS, Tastsoglou S, Kanellos I, et al. DIANA-TarBase v8: a decade-long collection of experimentally supported miRNA-gene interactions. Nucleic Acids Res. 2018;46:D239–45. https://doi.org/10.1093/nar/gkx1141.
Chou CH, Chang NW, Shrestha S, Da HS, Lin YL, Lee WH, et al. miRTarBase 2016: updates to the experimentally validated miRNA-target interactions database. Nucleic Acids Res. 2016;44:D239–47. https://doi.org/10.1093/nar/gkv1258.
Visualizer DS v4. 0. 100. 13345. Accelrys Sof tware Inc. 2005.
Piñero J, Bravo Á, Queralt-Rosinach N, Gutiérrez-Sacristán A, Deu-Pons J, Centeno E, et al. DisGeNET: a comprehensive platform integrating information on human disease-associated genes and variants. Nucleic Acids Res. 2017;45:D833–9. https://doi.org/10.1093/nar/gkw943.
Piñero J, Ramírez-Anguita JM, Saüch-Pitarch J, Ronzano F, Centeno E, Sanz F, et al. The DisGeNET knowledge platform for disease genomics: 2019 update. Nucleic Acids Res. 2020;48:D845–55. https://doi.org/10.1093/nar/gkz1021.
Lamb J, Crawford ED, Peck D, Modell JW, Blat IC, Wrobel MJ, et al. The connectivity map: using gene-expression signatures to connect small molecules, genes, and disease. Science (80- ). 2006;313:1929–35. https://doi.org/10.1126/science.1132939.
Berman HM, Westbrook J, Feng Z, Gilliland G, Bhat TN, Weissig H, et al. The protein data bank. Nucleic Acids Res. 2000:235–242. https://doi.org/10.1093/nar/28.1.235.
Waterhouse A, Bertoni M, Bienert S, Studer G, Tauriello G, Gumienny R, et al. SWISS-MODEL: homology modelling of protein structures and complexes. Nucleic Acids Res. 2018;46:W296–303. https://doi.org/10.1093/nar/gky427.
Kim S, Chen J, Cheng T, Gindulyte A, He J, He S, et al. PubChem 2019 update: improved access to chemical data. Nucleic Acids Res. 2019;47:D1102–9. https://doi.org/10.1093/nar/gky1033.
Morris GM, Ruth H, Lindstrom W, Sanner MF, Belew RK, Goodsell DS, et al. Software news and updates AutoDock4 and AutoDockTools4: automated docking with selective receptor flexibility. J Comput Chem. 2009;30:2785–91. https://doi.org/10.1002/jcc.21256.
Halgren TA. Merck molecular force field. I. Basis, form, scope, parameterization, and performance of MMFF94. J Comput Chem. 1996;17:490–519. https://doi.org/10.1002/(SICI)1096-987X(199604)17:5/6<490::AID-JCC1>3.0.CO;2-P.
Steffen C, Thomas K, Huniar U, Hellweg A, Rubner O, Schroer A. Software news and updates TmoleX-a graphical user interface for TURBOMOLE. J Comput Chem. 2010;31:2967–70. https://doi.org/10.1002/jcc.21576.
Trott O, Olson AJ. AutoDock Vina: improving the speed and accuracy of docking with a new scoring function, efficient optimization, and multithreading. J Comput Chem. 2010;31:455–61. https://doi.org/10.1002/jcc.21334.
Krieger, Elmar GV, Spronk C. YASARA - Yet Another Scientific Artificial Reality Application. YASARA.org. 2013.
Dickson CJ, Madej BD, Skjevik ÅA, Betz RM, Teigen K, Gould IR, et al. Lipid14: the amber lipid force field. J Chem Theory Comput. 2014;10. https://doi.org/10.1021/ct4010307.
Jorgensen WL, Chandrasekhar J, Madura JD, Impey RW, Klein ML. Comparison of simple potential functions for simulating liquid water. J Chem Phys. 1983;79. https://doi.org/10.1063/1.445869.
Krieger E, Nielsen JE, Spronk CAEM, Vriend G. Fast empirical pKa prediction by Ewald summation. J Mol Graph Model. 2006;25. https://doi.org/10.1016/j.jmgm.2006.02.009.
Krieger E, Vriend G. New ways to boost molecular dynamics simulations. J Comput Chem. 2015;36. https://doi.org/10.1002/jcc.23899.
Hess B, Bekker H, Berendsen HJC, Fraaije JGEM. LINCS: A Linear Constraint Solver for molecular simulations. J Comput Chem. 1997;18. https://doi.org/10.1002/(SICI)1096-987X(199709)18:12<1463::AID-JCC4>3.0.CO;2-H.
Miyamoto S, Kollman PA. Settle: An analytical version of the SHAKE and RATTLE algorithm for rigid water models. J Comput Chem. 1992;13. https://doi.org/10.1002/jcc.540130805.
Berendsen HJC, Postma JPM, Van Gunsteren WF, Dinola A, Haak JR. Molecular dynamics with coupling to an external bath. J Chem Phys. 1984;81. https://doi.org/10.1063/1.448118.
Mitra S, Dash R. Structural dynamics and quantum mechanical aspects of shikonin derivatives as CREBBP bromodomain inhibitors. J Mol Graph Model. 2018;83. https://doi.org/10.1016/j.jmgm.2018.04.014.
Srinivasan E, Rajasekaran R. Computational investigation of curcumin, a natural polyphenol that inhibits the destabilization and the aggregation of human SOD1 mutant (Ala4Val). RSC Adv. 2016;6. https://doi.org/10.1039/c6ra21927f.
Meriney SD, Fanselow EE. Synaptic Transmission. Synaptic Transmission. Elsevier; 2019. https://doi.org/10.1016/C2017-0-02762-3.
Gan L, Cookson MR, Petrucelli L, La Spada AR. Converging pathways in neurodegeneration, from genetics to mechanisms. Nature Neuroscience. 2018. pp. 1300–1309. https://doi.org/10.1038/s41593-018-0237-7.
Fong TG, Davis D, Growdon ME, Albuquerque A, Inouye SK. The interface between delirium and dementia in elderly adults. The Lancet Neurology. 2015. pp. 823–832. https://doi.org/10.1016/S1474-4422(15)00101-5.
Cerejeira J, Firmino H, Vaz-Serra A, Mukaetova-Ladinska EB. The neuroinflammatory hypothesis of delirium. Acta Neuropathol. 2010;119:737–54. https://doi.org/10.1007/s00401-010-0674-1.
Van Munster BC, Aronica E, Zwinderman AH, Eikelenboom P, Cunningham C, De Rooij SEJA. Neuroinflammation in Delirium: A Postmortem Case-Control Study. Rejuvenation Res. 2011;14:615. https://doi.org/10.1089/REJ.2011.1185.
Numakawa T, Odaka H. Brain-derived neurotrophic factor signaling in the pathophysiology of alzheimer’s disease: Beneficial effects of flavonoids for neuroprotection. International Journal of Molecular Sciences. 2021. p. 5719. https://doi.org/10.3390/ijms22115719.
Khansari N, Shakiba Y, Mahmoudi M. Chronic inflammation and oxidative stress as a major cause of age-related diseases and cancer. Recent Pat Inflamm Allergy Drug Discov. 2009;3:73–80. https://doi.org/10.2174/187221309787158371.
Reiner A, Levitz J. Glutamatergic Signaling in the Central Nervous System: Ionotropic and Metabotropic Receptors in Concert. Neuron. 2018. pp. 1080–1098. https://doi.org/10.1016/j.neuron.2018.05.018.
Praticò C, Quattrone D, Lucanto T, Amato A, Penna O, Roscitano C, et al. Drugs of anesthesia acting on central cholinergic system may cause post-operative cognitive dysfunction and delirium. Med Hypotheses. 2005;65:972–82. https://doi.org/10.1016/j.mehy.2005.05.037.
Robinson TN, Eiseman B. Postoperative delirium in the elderly: diagnosis and management. Clin Interv Aging. 2008;3:351–5. https://doi.org/10.2147/cia.s2759.
Janjua MS, Spurling BC, Arthur ME. Postoperative Delirium. StatPearls. Cham: Springer International Publishing; 2023. https://doi.org/10.1007/978-3-319-47771-8_23.
El Sheikh WG, Sleem B, Kobeissy F, Bizri M. Biomarkers of delirium and relation to dementia among the elderly in the intensive care unit: A narrative review. Biomarkers in Neuropsychiatry. 2023;8: 100064. https://doi.org/10.1016/j.bionps.2023.100064.
Baranyi A, Rothenhäusler HB. The impact of intra- and postoperative albumin levels as a biomarker of delirium after cardiopulmonary bypass: Results of an exploratory study. Psychiatry Res. 2012;200:957–63. https://doi.org/10.1016/j.psychres.2012.05.030.
Zhang DF, Su X, Meng ZT, Cui F, Li HL, Wang DX, et al. Preoperative severe hypoalbuminemia is associated with an increased risk of postoperative delirium in elderly patients: Results of a secondary analysis. J Crit Care. 2018;44:45–50. https://doi.org/10.1016/j.jcrc.2017.09.182.
Okawa Y, Tanaka K, Nakanishi Y, Asano T, Noji T, Ebihara Y, et al. The assessment of risk factors for postoperative delirium using cubic spline curves in gastroenterological surgery. Surg Today. 2021;51:1969–77. https://doi.org/10.1007/s00595-021-02379-2.
Kim Y, Jin Y, Jin T, Lee SM. Risk factors and outcomes of sepsis-associated delirium in intensive care unit patients: A secondary data analysis. Intensive Crit Care Nurs. 2020;59: 102844. https://doi.org/10.1016/j.iccn.2020.102844.
Diaz-Arrastia R, Wang KKW, Papa L, Sorani MD, Yue JK, Puccio AM, et al. Acute biomarkers of traumatic brain injury: Relationship between plasma levels of ubiquitin C-terminal hydrolase-l1 and glial fibrillary acidic protein. J Neurotrauma. 2014;31:19–25. https://doi.org/10.1089/neu.2013.3040.
Adamis D, Williams J, Finn K, Melvin V, Meagher D, McCarthy G. Brain-derived Neurotrophic Factor (BDNF) Levels and Delirium. Eur Psychiatry. 2017;41:s237–s237. https://doi.org/10.1016/j.eurpsy.2017.01.2263.
Williams J, Finn K, Melvin V, Meagher D, McCarthy G, Adamis D. The Association of Serum Levels of Brain-Derived Neurotrophic Factor with the Occurrence of and Recovery from Delirium in Older Medical Inpatients. Biomed Res Int. 2017;2017:5271395. https://doi.org/10.1155/2017/5271395.
Wyrobek J, Laflam A, Max L, Tian J, Neufeld KJ, Kebaish KM, et al. Association of intraoperative changes in brain-derived neurotrophic factor and postoperative delirium in older adults. Br J Anaesth. 2017;119:324–32. https://doi.org/10.1093/bja/aex103.
Siuda J, Patalong-Ogiewa M, Żmuda W, Targosz-Gajniak M, Niewiadomska E, Matuszek I, et al. Cognitive impairment and BDNF serum levels. Neurol Neurochir Pol. 2017;51:24–32. https://doi.org/10.1016/j.pjnns.2016.10.001.
Chen C, Li X, Wang T, Wang HH, Fu Y, Zhang L, et al. Association between NMDA receptor subunit 2b gene polymorphism and Alzheimer’s disease in Chinese Han population in Shanghai. Neurosci Bull. 2010;26:395–400. https://doi.org/10.1007/s12264-010-0729-2.
Jiang H, Jia J. Association between NR2B subunit gene (GRIN2B) promoter polymorphisms and sporadic Alzheimer’s disease in the North Chinese population. Neurosci Lett. 2009;450:356–60. https://doi.org/10.1016/j.neulet.2008.10.075.
Andreoli V, De Marco EV, Trecroci F, Cittadella R, Di Palma G, Gambardella A. Potential involvement of GRIN2B encoding the NMDA receptor subunit NR2B in the spectrum of Alzheimer’s disease. J Neural Transm. 2014;121:533–42. https://doi.org/10.1007/s00702-013-1125-7.
Han XR, Wen X, Wang YJ, Wang S, Shen M, Zhang ZF, et al. Effects of CREB1 gene silencing on cognitive dysfunction by mediating PKA-CREB signaling pathway in mice with vascular dementia. Mol Med. 2018;24:18. https://doi.org/10.1186/s10020-018-0020-y.
Wang H, Wang H, Song Y, Liu C, Qian X, Zhang D, et al. Overexpression of Foxc1 ameliorates sepsis-associated encephalopathy by inhibiting microglial migration and neuroinflammation through the IκBα/NF-κB pathway. Mol Med Rep. 2022;25:107. https://doi.org/10.3892/mmr.2022.12623.
Calsolaro V, Edison P. Neuroinflammation in Alzheimer’s disease: Current evidence and future directions. Alzheimer’s and Dementia. Elsevier Inc.; 2016. pp. 719–732. https://doi.org/10.1016/j.jalz.2016.02.010.
Maiese K. Forkhead transcription factors: new considerations for alzheimer’s disease and dementia. J Transl Sci. 2016;2:241–7. https://doi.org/10.15761/jts.1000146.
Tam KT, Chan PK, Zhang W, Law PP, Tian Z, Chan GCF, et al. Identification of a novel distal regulatory element of the human Neuroglobin gene by the chromosome conformation capture approach. Nucleic Acids Res. 2017;45:115–26. https://doi.org/10.1093/nar/gkw820.
Nakada C, Satoh S, Tabata Y, Arai K, Watanabe S. Transcriptional Repressor foxl1 Regulates Central Nervous System Development by Suppressing shh Expression in Zebra Fish. Mol Cell Biol. 2006;26:7246–57. https://doi.org/10.1128/mcb.00429-06.
Rahman MR, Islam T, Turanli B, Zaman T, Faruquee HM, Rahman MM, et al. Network-based approach to identify molecular signatures and therapeutic agents in Alzheimer’s disease. Comput Biol Chem. 2019;78:431–9. https://doi.org/10.1016/j.compbiolchem.2018.12.011.
Turk A, Kunej T, Peterlin B. Microrna-target interaction regulatory network in alzheimer’s disease. J Pers Med. 2021;11:1275. https://doi.org/10.3390/jpm11121275.
Kim Y-J, Kim SH, Park Y, Park J, Lee JH, Kim BC, et al. miR-16-5p is upregulated by amyloid β deposition in Alzheimer’s disease models and induces neuronal cell apoptosis through direct targeting and suppression of BCL-2. Exp Gerontol. 2020;136: 110954. https://doi.org/10.1016/j.exger.2020.110954.
Coccia E, Masanas M, López-Soriano J, Segura MF, Comella JX, Pérez-García MJ. FAIM Is Regulated by MiR-206, MiR-1–3p and MiR-133b. Front Cell Dev Biol. 2020;8. https://doi.org/10.3389/fcell.2020.584606.
Sarkar S, Engler-Chiurazzi EB, Cavendish JZ, Povroznik JM, Russell AE, Quintana DD, et al. Over-expression of miR-34a induces rapid cognitive impairment and Alzheimer’s disease-like pathology. Brain Res. 2019;1721: 146327. https://doi.org/10.1016/j.brainres.2019.146327.
Cosín-Tomás M, Antonell A, Lladó A, Alcolea D, Fortea J, Ezquerra M, et al. Plasma miR-34a-5p and miR-545-3p as Early Biomarkers of Alzheimer’s Disease: Potential and Limitations. Mol Neurobiol. 2017;54:5550–62. https://doi.org/10.1007/s12035-016-0088-8.
Li P, Xu Y, Wang B, Huang J, Li Q. miR-34a-5p and miR-125b-5p attenuate Aβ-induced neurotoxicity through targeting BACE1. J Neurol Sci. 2020;413: 116793. https://doi.org/10.1016/j.jns.2020.116793.
Hu YT, Chen XL, Huang SH, Zhu Q Bin, Yu SY, Shen Y, et al. Early growth response-1 regulates acetylcholinesterase and its relation with the course of Alzheimer’s disease. Brain Pathol. 2019;29: 502–512. https://doi.org/10.1111/bpa.12688.
Citron BA, Saykally JN, Cao C, Dennis JS, Runfeldt M, Arendash GW. Transcription factor Sp1 inhibition, memory, and cytokines in a mouse model of Alzheimer’s disease. Am J Neurodegener Dis. 2015;4: 40–8. Available: /pmc/articles/PMC4700125/.
Citron BA, Dennis JS, Zeitlin RS, Echeverria V. Transcription factor Sp1 dysregulation in Alzheimer’s disease. J Neurosci Res. 2008;86:2499–504. https://doi.org/10.1002/jnr.21695.
Wu HZY, Ong KL, Seeher K, Armstrong NJ, Thalamuthu A, Brodaty H, et al. Circulating microRNAs as biomarkers of alzheimer’s disease: A systematic review. Hornberger M, editor. J Alzheimer’s Dis. 2015;49: 755–766. https://doi.org/10.3233/JAD-150619.
Abuelezz NZ, Nasr FE, AbdulKader MA, Bassiouny AR, Zaky A. MicroRNAs as Potential Orchestrators of Alzheimer’s Disease-Related Pathologies: Insights on Current Status and Future Possibilities. Frontiers in Aging Neuroscience. Frontiers Media SA; 2021. p. 743573. https://doi.org/10.3389/fnagi.2021.743573.
Mainerova B, Praskoa J, Latalova K, Axmann K, Cerna M, Horacek R, et al. Alcohol withdrawal delirium - diagnosis, course and treatment. Biomed Pap. 2015;159:44–52. https://doi.org/10.5507/bp.2013.089.
Blondell RD. Trauma and Alcohol. Comprehensive Handbook of Alcohol Related Pathology. Elsevier; 2005. pp. 321–335. https://doi.org/10.1016/B978-012564370-2/50028-3.
Gill TS. The delirious patient. Handb Emerg Psychiatry. Elsevier; 2007:106–120. https://doi.org/10.1016/b978-0-323-04088-4.50011-8.
Crismon ML. Tacrine: First drug approved for alzheimer’s disease. Ann Pharmacother. 1994;28:744–51. https://doi.org/10.1177/106002809402800612.
Knapp MJ, Knopman DS, Solomon PR, Pendlebury WW, Davis CS, Gracon SI. A 30-week randomized controlled trial of high-dose tacrine in patients with Alzheimer’s disease. JAMA J Am Med Assoc. 1994;271:985–91. https://doi.org/10.1001/jama.1994.03510370037029.
Watkins PB, Zimmerman HJ, Knapp MJ, Gracon SI, Lewis KW. Hepatotoxic effects of tacrine administration in patients With Alzheimer’s disease. JAMA J Am Med Assoc. 1994;271:992–8. https://doi.org/10.1001/jama.1994.03510370044030.
Romero A, Cacabelos R, Oset-Gasque MJ, Samadi A, Marco-Contelles J. Novel tacrine-related drugs as potential candidates for the treatment of Alzheimer’s disease. Bioorg Med Chem Lett. 2013:1916–1922. https://doi.org/10.1016/j.bmcl.2013.02.017.
Dawson AH, Buckley NA. Pharmacological management of anticholinergic delirium - theory, evidence and practice. Br J Clin Pharmacol. 2016. https://doi.org/10.1111/bcp.12839.
Ongnok B, Khuanjing T, Chunchai T, Kerdphoo S, Jaiwongkam T, Chattipakorn N, et al. Donepezil provides neuroprotective effects against brain injury and Alzheimer’s pathology under conditions of cardiac ischemia/reperfusion injury. Biochim Biophys Acta - Mol Basis Dis. 2021;1867:165975. https://doi.org/10.1016/j.bbadis.2020.165975.
Wirsich J, Rey M, Guye M, Bénar C, Lanteaume L, Ridley B, et al. Brain networks are independently modulated by donepezil, sleep, and sleep deprivation. Brain Topogr. 2018;31:380–91. https://doi.org/10.1007/s10548-017-0608-5.
Mishra CB, Kumari S, Manral A, Prakash A, Saini V, Lynn AM, et al. Design, synthesis, in-silico and biological evaluation of novel donepezil derivatives as multi-target-directed ligands for the treatment of Alzheimer’s disease. Eur J Med Chem. 2017;125:736–50. https://doi.org/10.1016/j.ejmech.2016.09.057.
Junaid M, Islam N, Hossain MK, Ullah MO, Halim MA. Metal based donepezil analogues designed to inhibit human acetylcholinesterase for Alzheimer’s disease. Silman I, editor. PLoS One. 2019;14:e0211935. https://doi.org/10.1371/journal.pone.0211935.
Khosravan A, Marani S, Sadeghi Googheri MS. The effects of fluorine substitution on the chemical properties and inhibitory capacity of donepezil anti-Alzheimer drug; density functional theory and molecular docking calculations. J Mol Graph Model. 2017;71:124–34. https://doi.org/10.1016/j.jmgm.2016.11.013.
Hefner G, Brueckner A, Hiemke C, Fellgiebel A. Therapeutic drug monitoring for patients with alzheimer dementia to improve treatment with donepezil. Ther Drug Monit. 2015;37:353–61. https://doi.org/10.1097/FTD.0000000000000152.
Sampson EL, Raven PR, Ndhlovu PN, Vallance A, Garlick N, Watts J, et al. A randomized, double-blind, placebo-controlled trial of donepezil hydrochloride (Aricept) for reducing the incidence of postoperative delirium after elective total hip replacement. Int J Geriatr Psychiatry. 2007;22. https://doi.org/10.1002/gps.1679.
Liptzin B, Laki A, Garb JL, Fingeroth R, Krushell R. Donepezil in the prevention and treatment of post-surgical delirium. Am J Geriatr Psychiatry. 2005;13. https://doi.org/10.1097/00019442-200512000-00010.
Lieberman OJ, Lee S, Zabinski J. Donepezil treatment is associated with improved outcomes in critically ill dementia patients via a reduction in delirium. Alzheimer’s Dement. 2023;19:1742–51. https://doi.org/10.1002/alz.12807.
Lovering AL, Seung SL, Kim YW, Withers SG, Strynadka NCJ. Mechanistic and structural analysis of a family 31 α-glycosidase and its glycosyl-enzyme intermediate. J Biol Chem. 2005;280. https://doi.org/10.1074/jbc.M410468200.
Acknowledgements
The authors are acknowledging the reviewer’s contribution to assist in the study screening and selection process, their comments, and suggestions for improving the quality of the manuscript.
Funding
This work formed part of the first author’s PhD research at the University of Southern Queensland, Australia which has been supported by funding from the Australian Government Research Training Program Scholarship.
Author information
Authors and Affiliations
Contributions
MPM conceptualized, collected data, analyzed, and wrote the first draft of the study. KA, JG, NHM, and RAM supervised the study as well as revised and edited the manuscript.
Corresponding author
Ethics declarations
Ethics approval and consent to participate
Not applicable.
Consent for publication
Not applicable.
Competing interests
The authors declare no competing interests.
Additional information
Publisher’s Note
Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.
Rights and permissions
Open Access This article is licensed under a Creative Commons Attribution-NonCommercial-NoDerivatives 4.0 International License, which permits any non-commercial use, sharing, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if you modified the licensed material. You do not have permission under this licence to share adapted material derived from this article or parts of it. The images or other third party material in this article are included in the article’s Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article’s Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://creativecommons.org/licenses/by-nc-nd/4.0/.
About this article
Cite this article
Mosharaf, M.P., Alam, K., Gow, J. et al. Common molecular and pathophysiological underpinnings of delirium and Alzheimer’s disease: molecular signatures and therapeutic indications. BMC Geriatr 24, 716 (2024). https://doi.org/10.1186/s12877-024-05289-3
Received:
Accepted:
Published:
DOI: https://doi.org/10.1186/s12877-024-05289-3