Open Access

Preliminary exploration of the expression of acetylcholinesterase in normal human T lymphocytes and leukemic Jurkat T cells

  • Authors:
    • José Luis Gómez‑Olivares
    • Rosa María López‑Durán
    • Sergio Enríquez‑Flores
    • Gabriel López‑Velázquez
    • Ignacio De La Mora‑De La Mora
    • Itzhel García‑Torres
    • Rubí Viedma‑Rodríguez
    • Rafael Valencia‑Quintana
    • Mirta Milić
    • Luis Antonio Flores‑López
  • View Affiliations

  • Published online on: August 28, 2024     https://doi.org/10.3892/br.2024.1846
  • Article Number: 158
  • Copyright: © Gómez‑Olivares et al. This is an open access article distributed under the terms of Creative Commons Attribution License.

Metrics: Total Views: 0 (Spandidos Publications: | PMC Statistics: )
Total PDF Downloads: 0 (Spandidos Publications: | PMC Statistics: )


Abstract

The classic enzymatic function of acetylcholinesterase (AChE) is the hydrolysis of acetylcholine (ACh) in the neuronal synapse. However, AChE is also present in nonneuronal cells such as lymphocytes. Various studies have proposed the participation of AChE in the development of cancer. The ACHE gene produces three mRNAs (T, H and R). AChE‑T encodes amphiphilic monomers, dimers, tetramers (G1A, G2A and G4A) and hydrophilic tetramers (G4H). AChE‑H encodes amphiphilic monomers and dimers (G1A and G2A). AChE‑R encodes a hydrophilic monomer (G1H). The present study considered the differences in the mRNA expression (T, H and R) and protein levels of AChE, as well as the molecular forms of AChE, the glycosylation pattern and the enzymatic activity of AChE present in normal T lymphocytes and leukemic Jurkat E6‑1 cells. The results revealed that AChE enzymatic activity was higher in normal T lymphocytes than in Jurkat cells. Normal T cells expressed AChE‑H transcripts, whereas Jurkat cells expressed AChE‑H and AChE‑T. The molecular forms identified in normal T cells were G2A (5.2 S) and G1A (3.5 S), whereas those in Jurkat cells were G2A (5.2 S), G1A (3.5 S) and G4H (10.6S). AChE in Jurkat cells showed altered posttranslational maturation since a decrease in the incorporation of galactose and sialic acid into its structure was observed. In conclusion, the content and composition of AChE were altered in Jurkat cells compared with those in normal T lymphocytes. The present study opened new avenues for exploring the development of novel therapeutic strategies against T‑cell leukemia and for identifying potential molecular targets for the early detection of this type of cancer.

Introduction

The classical enzymatic function of AChE is the precise temporary hydrolysis of ACh at the cholinergic synapse. However, the biological role of AChE is not limited to cholinergic transmission (1,2). The aforementioned findings are supported by the high levels of AChE activity estimated in nonneuronal tissues, such as blood cells, erythrocytes, lymphocytes and megakaryocytes (3). The structure and functions of AChE have been the subject of various investigations worldwide. However, a number of unresolved questions remain, both in terms of its mechanism of action, synthesis processes and transport through cellular compartments. The physiological meaning of AChE molecular polymorphisms is also unknown, although their roles in apoptosis, proliferation, cell differentiation, cell adhesion, neurogenesis and hematopoiesis and possible participation in other biological activities have begun to be elucidated (1,4-8).

The structural complexity of human AChE is due to a gene with a size of 7 kb located in the q22 region of chromosome 7 (9,10). The ACHE gene produces three different mRNAs (T, H and R) through an alternative splicing process (11). The generated mRNAs share exons E2, E3 and E4, which encode the catalytic domain of AChE (12,13). The AChE-T-type mRNA is assembled with exons E2-E3-E4 and E6; it encodes amphiphilic monomers, dimers and tetramers (G1A, G2A and G4A), hydrophilic tetramers (G4H) and hetero-oligomeric associations with a collagen-like stem Q, structurally conforming to the asymmetric forms (A4, A8 and A12, which contain one, two and three tetramers, respectively), or with the transmembrane protein proline-rich membrane anchor (PRiMA), constituting membrane-bound tetramers PRiMA-G4A (14-16). The AChE-H-type mRNA contains exons E2-E3-E4-E5, which encode Type I amphiphilic monomers and dimers (G1A and G2A) in which each subunit has a covalently linked glycophosphatidylinositol residue (11,16,17). The AChE-R-type mRNA, also called read-through, does not undergo splicing after the last exon encoding the catalytic domain and, therefore, has the arrangement E2-E3-E4-I4-E5. The C-terminus of the R subunit consists of 30 amino acids and lacks Cys and thus the subunits remain as hydrophilic monomers (G1H) (18,19). Abnormalities in the ACHE gene have been observed in various types of cancer, such as ovarian, breast, prostate, liver and some types of myeloid leukemia (5,20-22).

In general, both brain and nonbrain tumors present an alteration in AChE activity; in a number of cases, a decrease in specific AChE activity is observed (23,24). Lung cancer shows a decrease in AChE-T mRNA levels (25), whereas colon cancer shows decreases in AChE-T, AChE-H and AChE-R mRNA levels (26,27). The decrease in mRNA levels is related to a decrease in AChE activity (25,28). Furthermore, alterations in the ACHE gene in breast cancer decrease AChE enzymatic activity in axillary lymph nodes (29). In other types of cancer, such as in gliomas (30), liver tumors (5,31), gastric cancer (32) and head and neck carcinomas, a reduction in the enzymatic activity of AChE has been observed, where the low activity of AChE is associated with worse survival (33,34). In patients with prostate cancer, a decrease in AChE activity in the blood has been reported compared with that in a control group of healthy subjects; this finding opens the possibility of proposing AChE as a cancer biomarker (35). All these reported events, where a decrease in both the expression and enzymatic activity of AChE occur, cause an increase in the amount of ACh that leads to cholinergic overstimulation and an increase in cell proliferation (36-40).

Antisense polynucleotides against AChE produce essential changes in hematopoiesis, decreasing apoptosis and increasing the proliferation of blood cells (41,42). Organophosphate pesticides are potent inhibitors of AChE and their use in agricultural activities is associated with the development of non-Hodgkin lymphoma and leukemia (21,43-46). Other studies of rat mammary glands have demonstrated that the inhibition of AChE with physostigmine induces cancer development, revealing a relationship between a decrease in AChE activity and carcinogenesis (47).

A clear relationship exists between low AChE activity and a poor prognosis for cancer patients. However, determining why reducing AChE activity and protein content induces tumor progression and aggressiveness is important. With respect to the information available regarding AChE gene and protein expression in cancerous tissues, research still needs to be performed to clarify the molecular mechanisms by which AChE could participate in types of cancer. Exploring study models that reveal new information about the relationship of this enzyme with the processes of altered cell proliferation is essential (5). Based on the aforementioned findings, hematopoietic cells, mainly T lymphocytes, have a well-established nonneuronal cholinergic system called the lymphocytic cholinergic system (3,48-50), which is why the present study proposed it as a study model to obtain an improved understanding of the differences in AChE expression between normal cells and cancer cells.

Based on the reported evidence that AChE expression is altered in cancer cells compared with that in their normal counterparts, the present study aimed to compare AChE expression at the mRNA and protein levels, as well as the glycosylation and enzymatic activity of the AChE protein between normal T lymphocytes and the human T lymphoblast cell line Jurkat E6-1.

Materials and methods

Reagents and general materials

Fetal bovine serum (FBS), penicillin, streptomycin, L-glutamine and sodium pyruvate solutions were purchased from Invitrogen (Thermo Fisher Scientific, Inc.). The other reagents mentioned were acquired from MilliporeSigma. ethylenediaminetetraacetic acid (EDTA)-Vacutainer tubes for blood collection from healthy volunteers were acquired from Becton, Dickinson and Company.

Cell culture

The human T lymphoblast cell line Jurkat E6-1 (TIB-152) was obtained from the American Type Culture Collection. Jurkat E6-1 cells were cultured in RPMI-1640 medium (MilliporeSigma) supplemented with 10% (v/v) FBS, 2 mM L-glutamine, 1 mM sodium pyruvate, 100 U/ml penicillin and 100 mg/ml streptomycin. Jurkat cells from passages 2 to 5 were grown at 37˚C in a humidified 5% CO2 atmosphere for all experiments.

The human liver cancer cell line HepG2 (HB-8065) was obtained from the American Type Culture Collection. HepG2 cells were cultured in DMEM (MilliporeSigma) supplemented with 10% (v/v) FBS, 2 mM L-glutamine, 1 mM sodium pyruvate, 100 U/ml penicillin and 100 mg/ml streptomycin at 37˚C in a humidified atmosphere containing 5% CO2.

Obtaining T lymphocytes from healthy adults

The use of lymphocytes from healthy donors was included to perform comparative tests with leukemic cells. The sample consisted of 10 young adults (aged 18-22 years) selected from the period between January, 2023 and April, 2023, according to the following inclusion and exclusion criteria.

The inclusion criteria were: Age between 18-22 years, weight between 50-80 kg, systolic blood pressure between 90-160 mm Hg, diastolic blood pressure between 60-90 mm Hg, pulse between 50-100 beats and fasting for 6-8 h. The volunteers could not consume foods with fat, eggs, milk, or their derivatives 24 h before donation.

The exclusion criteria were as follows: Pregnant or breastfeeding women; individuals who had suffered from flu, cough, diarrhea or dental infection in the last 14 days; individuals who had taken medications in the last five days; individuals who had undergone endodontic treatment, acupuncture or had tattoos or piercings in the last 12 months; those who had undergone surgery in the last six months; those who had been vaccinated in the last 30 days; and those who consumed alcoholic beverages in the 72 h prior to donation.

Informed consent was obtained from all individual participants included in the study to the collection and use of their cells strictly for research purposes. The research ethics committee approved the use of blood samples donated by healthy adults with a registration in the National Bioethics Commission with the number CONBIOETICA-09-CEI-025-20161215 of the National Institute of Pediatrics, Mexico.

Density gradient separation was performed from 20 ml of blood, which was generated via density gradient separation using Lymphoprep density gradient medium (STEMCELL Technologies) and centrifuged at 800 x g for 20 min at room temperature (20-25˚C) to extract the mononuclear cells (including T lymphocytes), which were subsequently resuspended in 1 ml of PBS.

Magnetic beads conjugated with an anti-CD3 antibody (specific T lymphocyte marker) MAC (Miltenyi Biotec, Inc.) were added to the suspension of mononuclear cells. T lymphocytes were isolated by magnetic separation with MiniMACS Starting Kit (Miltenyi Biotec, Inc.) according to the manufacturer's protocol. This device collects the entire cell suspension and separates the cells into a positive fraction (T lymphocytes) and a negative fraction (the remainder of the cells). The magnetic field of the MACS column allows for minimal labeling of target cells with small microbeads. This means that there are plenty of surface epitopes available for fluorescent staining and flow cytometry analysis. The positive fraction was recovered and the purity of the T lymphocytes was determined via flow cytometry with a FACSCalibur 2 Laser, 4 Color flow cytometer (Becton, Dickinson and Company); the CD3-PerCP antibody was used incubated for 30 min at room temperature (20-25˚C). The samples were processed with CellQuest Pro v 5.2.1. Software (BD Biosciences) via the SSC and the association with CD3-PerCP.

Trypan blue (MicroLab-Mexico) was used to determine the viability and density of total leukocytes, purified T lymphocytes and Jurkat cells. Cells were incubated with equal volume of 0.4% trypan blue solution for 3 min at room temperature (20-25˚C). Viable cells at a density of 1x107 cells/ml were used for the assays.

Total RNA extraction

Total RNA was extracted from 10x107 cell samples with 1 ml of TRIzol (Thermo Fisher Scientific, Inc.) according to the manufacturer's instructions. The isolated RNA was partially dissolved in 100 µl of RNase-free water and incubated at 55˚C for 5 min. The RNA concentration and purity were determined with a Nanodrop spectrophotometer (Thermo Fisher Scientific, Inc.) by calculating the optical density ratio at wavelengths of 260/280 nm and 260/230 nm.

Reverse transcription (RT)-PCR amplification

The RT-PCR process was performed using the one-step RT-PCR kit (Qiagen, Inc.) and the following primer pairs: AChE-H (487 bp), forward 5'-TCTCGAAACTACACGGCAGA-3' and reverse 5'-TGAGGAGGAAGGGAGCACTA-3'; AChE-T (444 bp), forward 5'-TCTCGAAACTACACGGCAGA-3' and reverse 5'-GCCCAGCCCTGAAATAAATAG-3'; AChE-R (333 bp), forward 5'-TCTCGAAACTACACGGCAGA-3' and reverse 5'-GGGGAGAAGAGAGGGGTTAC-3'; and β-actin (500 bp) as a housekeeping gene, forward 5'-CACTGGCATCGTGATGGACT-3' and reverse 5'-AATGCCAGGGTACATGGTGG-3'. A total of 1 µM of each primer, 1 µl of a 40 mM dNTP mixture, 5 µl of 5X Q solution, 1 µl of enzyme mix and 1.5 µg of RNA sample were used (titrations were performed using different amounts of RNA, 0.5-2.0 µg), for a final volume of 25 µl that was completed with RNase-free water.

The samples were placed in a Mastercycler Gradient Thermal Cycler (Eppendorf SE) and cycled with the following program: 50˚C for 30 min and 95˚C for 15 min for reverse transcription, followed by 39 cycles (1 min at 94˚C, 1 min at 50˚C and 1 min at 72˚C) for amplification. After the final cycle, the temperature was maintained at 72˚C for 10 min. The amplified DNA fragments were visualized on a 2% (w/v) agarose gel containing 0.5 µg/ml ethidium bromide. Image acquisition was performed using a Molecular Imager Gel Doc XR + system (Bio-Rad Laboratories, Inc.). The optical density of each band was calculated after background subtraction and normalization to β-actin for semiquantitative analysis using Image Studio 4.0 software (LI-COR Biosciences).

Solubilization of the AChE protein

For AChE solubilization from isolated T lymphocytes and Jurkat cells, 1x107 cells were used per sample. Each sample was homogenized with HEPES saline buffer [15 mM HEPES, 1 M NaCl, 50 mM MgCl2, 1 mM ethylene glycol-bis(2-aminoethylether)-N,N,N',N'-tetraacetic acid and 3 mM EDTA, pH 7.5] with a mixture of protease inhibitors at 10% cell weight/buffer volume (w/v). The homogenate was centrifuged at 200,000 x g in a Beckman Ti 60 rotor (Beckman Coulter, Inc.) for 1 h at 4˚C to separate soluble AChE from AChE weakly bound to the membranes. After the first supernatants (S1) were collected, the precipitates were homogenized with HEPES saline buffer supplemented with 1% (w/v) Triton X-100. These suspensions were subsequently centrifuged at 200,000 x g for 1 h at 4˚C to obtain the enzyme bound to the membranes in the second supernatant (S2). Centrifugation was performed using a Beckman XL-90 Optima centrifuge (Beckman Coulter, Inc.). The enzymatic activity of AChE was subsequently determined in both fractions (S1 and S2).

Estimation of AChE enzymatic activity

The measurement of AChE activity was performed via spectrophotometry with a Cary 50 spectrophotometer (Agilent Technologies, Inc.) using the Ellman method (51). The reaction medium contained 100 mM phosphate buffer, pH 8.0, with 0.33 mM 5,5'-dithiobis-(2-nitrobenzoic) acid, 1 mM acetylthiocholine iodide and 50 µM Iso-OMPA, butyrylcholinesterase inhibitor. One AChE unit (U) represents the enzyme that hydrolyzes 1 µmol of substrate per min at 37˚C.

AChE activity in the fractions collected from the sucrose gradients was determined using the Ellman method adapted to a microassay method using a microplate spectrophotometer Epoch (BioTek; Agilent Technologies, Inc.), for which transparent plastic plates were used (Nunc), with 96 wells of 400 µl and a flat bottom. Enzymatic activity was expressed in arbitrary units (AUs) such that one AU represented an increase in A405 of 0.001 per min and for each µl of the sample at room temperature. The Bradford method was used to determine the protein content with bovine serum albumin as a standard protein (52).

Profile of the molecular forms of the AChE protein

The molecular forms of AChE were determined by gradient centrifugation using two sucrose solutions of different concentrations (5 and 20% w/v) prepared in 10 mM Tris-HCl buffer, pH 7.0, with 1 M NaCl, 50 mM MgCl2 and Triton X-100 (0.5%, W/V). The proteins used as sedimentation standards were bovine liver catalase (11.4S) and bovine intestine phosphatase (6.1S). The gradients were centrifuged at 200,000 x g for 18 h at 4˚C in an SW41Ti tilting rotor (Beckman Coulter, Inc.). Following centrifugation, 37-40 fractions were collected using the peristaltic pump Econo Gradient Pump and a fraction collector model 2110 (both from Bio-Rad Laboratories, Inc.). Sedimentation coefficients were determined according to Martin and Ames's method (53), comparing the distance traveled by the AChE protein with that of standard proteins.

AChE protein glycosylation profile

Lectins from Concanavalia ensiformis (ConA), Lens culinaris (LCA), Triticum vulgaris (WGA) and Ricinus communis (RCA) were used to determine possible differences in the binding of AChE from T lymphocytes and Jurkat cells to glycans. ConA recognizes mannose residues, LCA recognizes D-mannose bound to fucosylation centers, WGA recognizes N-acetyl-glucosamine and RCA recognizes galactose or sialic acid.

Aliquots of the mixture of S1+S2 extracts (0.5 ml) of T lymphocytes and Jurkat cells were incubated with 0.25 ml of Sepharose-4B (nonspecific binding enzyme control) or ConA, LCA, RCA or WGA overnight at 4˚C with constant stirring. AChE activity in the supernatants was assessed via the microassay method to calculate the percentage of protein interacting with lectin and the supernatant corresponding to Sepharose-4B, which represented 100% of the non-retained proteins, was used as a reference.

Statistical analysis

The values of the means and variances for solubilized enzyme activity, protein content, interaction with lectins and the proportion of each molecular form were obtained for normal lymphocytes and leukemic cells for statistical analysis. The results were compared using the Mann-Whitney U test to establish a significant difference. Data were analyzed using NCSS 2007 Statistical Software (NCSS, LLC). P<0.05 was considered to indicate a statistically significant difference.

Results

Collection of T lymphocytes and analysis of purity using flow cytometry

The average number of leukocytes isolated from 10 blood samples of 20 ml was 7.13x107±0.89. The average number of T lymphocytes isolated from the total leukocyte population by magnetic separation was 1.02x107±1.12 and their viability was 100% (Table I). Fig. 1 shows the magnetic separation of T lymphocytes by flow cytometry. Of the total leukocytes, ~25% were positive for the CD3 surface marker, indicating the presence of T lymphocytes. The average percentage of the 10 samples analyzed using flow cytometry was 98.36%, indicating that the cell purification process provided high performance.

Table I

Viability and density of total leukocytes and purified T lymphocytes.

Table I

Viability and density of total leukocytes and purified T lymphocytes.

SourceCell number/ml (n=10)Viability (n=10)
Leukocytes 7.13x107±0.89100%
T lymphocytes 1.02x107±1.12100%

[i] Peripheral blood samples of 20 ml each were obtained from 10 healthy adults. Total leukocytes were obtained by density gradient separation with Lymphoprep and T lymphocytes were purified by magnetic separation using an anti-CD3 antibody conjugated to magnetic beads. Cell viability and density were determined using Trypan blue. The table shows the averages and standard deviations of the number of cells/ml and the percentage of cell viability obtained at each stage of cell separation.

Jurkat cells were also labeled with CD3-PerCP to determine the percentage of CD3+ cells. The results revealed that 84.89% of these cells were positive for the CD3 surface marker. Similarly, Fig. 1C shows an analysis performed using flow cytometry on a sample of Jurkat cells labeled with CD3-PerCP.

One-step RT-PCR of the ACHE mRNA

Normal human T lymphocytes were used for mRNA titration with different primers. The results revealed that the optimal concentration for amplification assays was 1.5 µg of RNA.

The detection of AChE mRNA variants in normal T cells revealed only one amplified product, corresponding to the AChE-H transcript; the products of the amplifications revealed a band in the gel of ~487 bp (Fig. 2). These results indicated that human T lymphocytes express only the AChE-H transcript, which encodes membrane-anchored amphiphilic dimers (17,54).

In the analysis of the expression of the ACHE gene transcripts in Jurkat leukemia cells, the results revealed that the transcripts detected corresponded to AChE-H (487 bp) and AChE-T (444 bp); the latter was not detected in normal T lymphocytes (Fig. 3). With these data, a comparative analysis of the expression of AChE in T lymphocytes and leukemic cells could be later conducted. Fig. 4 shows the amplification products of the alternative AChE transcripts detected in normal T cells and Jurkat cells; as a positive control for AChE-R, a 574 bp product was obtained using the same set of primers for AChE-H and mRNA extracted from HepG2 cells.

Finally, densitometry analyses were performed to obtain the intensity of each band, thus obtaining an average of the intensity of the β-actin control, which facilitated establishing the relative intensity index of each band of the amplicons corresponding to AChE in each sample (Table II).

Table II

Relative intensity of mRNA splicing isoforms for the ACHE gene.

Table II

Relative intensity of mRNA splicing isoforms for the ACHE gene.

 Expressed AChE transcript
Cell typeHRT
Normal T lymphocytes0.2326±0.0102NDND
Jurkat E6-1 cells0.1909±0.0050ND0.4638±0.0058
HepG2 cells0.2615±0.03080.0657±0.0128-

[i] The values represent the means and respective standard deviations of four data points obtained from the densitometry analysis of samples of normal T lymphocytes (n=4), Jurkat cells (n=3) and HepG2 cells (n=3); the last was used as a positive control for the expression of the AChE-R transcript. The reverse transcription PCR data were normalized to the control of the β-actin gene expressed in each cell type, with a mean value of 206.39±3.5756. The expression of the AChE-H transcript was significantly higher in normal T cells than in Jurkat cells (Mann-Whitney U test; P<0.05). AChE, acetylcholinesterase.

Estimation of AChE activity in normal T lymphocytes and Jurkat cells

The AChE activity in the enzyme extracts of T lymphocytes solubilized by ionic force (S1) or detergent (S2) did not change. These results indicated that in human T lymphocytes, a similar proportion of AChE is weakly bound and that AChE is strongly bound to the membrane.

On the other hand, the results obtained from the solubilization of the AChE protein in Jurkat cells revealed higher AChE activity in the weakly membrane-bound fraction (S1) than in the strongly membrane-bound fraction (S2). Table III shows the protein content, AChE activity and AChE-specific activity obtained from the S1 and S2 extracts of normal T lymphocytes and Jurkat cells. Specifically, Jurkat cells had lower activity than normal T lymphocytes. These data were thought-provoking as, although the Jurkat cells expressed a transcript, their activity levels decreased.

Table III

Solubilized acetylcholinesterase activity of normal T lymphocytes and Jurkat cells.

Table III

Solubilized acetylcholinesterase activity of normal T lymphocytes and Jurkat cells.

 Protein (mg/ml)AChE (U/ml)A.E. AChE (U/mg)
T-lymphocytes   
     Fraction S11.96±0.00350.216±0.08490.1101±0.0433
     Fraction S22.22±0.00490.294±0.02550.1323±0.0115
Jurkat Cells   
     Fraction S14.86±0.00450.287±0.03120.0590±0.0064
     Fraction S22.50±0.01640.088±0.00000.0353±0.0000

[i] Means and standard deviations of the protein content, AChE activity by volume and A.E. estimated from solubilization of T-lymphocyte (n=3) and Jurkat cell (n=3) samples. Fraction S1 was saline HEPES buffer and S2 was saline HEPES buffer containing Triton X-100. The AChE activity in the S1 and S2 fractions of T lymphocytes was similar (Mann-Whitney U test, P≤0.05). In the case of Jurkat cells, the AChE activity in the S1 fraction was higher than that in the S2 fraction. The A.E. of AChE in the S1 and S2 fractions of Jurkat cells was lower than that in normal T cells. A.E., specific activity; AChE, acetylcholinesterase.

Glycosylation profile of the AChE protein in T lymphocytes and Jurkat cells

An analysis of the interaction with lectins revealed the posttranslational maturation process that glycoproteins undergo. In the case of T lymphocytes, the AChE protein had a weak interaction with the ConA lectin (33%), the interaction with LCA was 66%, the interaction level observed with WGA was 63% and the interaction with RCA was 75% (Fig. 5). Meanwhile, the glycosylation profile of the AChE protein differed in Jurkat leukemic cells. The interactions of the AChE protein with the ConA, LCA, WGA and RCA lectins were 66, 72, 67 and 67%, respectively. The difference in the level of AChE glycosylation between T lymphocytes and leukemic T cells indicated that the posttranslational maturation process of AChE is modified in the neoplastic state.

Profile of the molecular forms of AChE in T lymphocytes and Jurkat cells

The results for the molecular forms of AChE with sedimentation coefficients of 5.2S and 3.5S were consistent with previously reported values for amphiphilic dimers (G2A, 5.2S) and amphiphilic monomers (G1A, 3.5S) found in T-cell extracts (55,56). In the sedimentation analysis of the Jurkat cell extracts, amphiphilic dimers (G2A, 5.2S), amphiphilic monomers (G1A, 3.5S) and hydrophilic tetramers (G4H, 10.6) were the AChE molecular forms identified. The latter is encoded by the AChE-T transcript, thus confirming the presence of this alternative splicing variant in the studies conducted via RT-PCR (57). Fig. 6 shows the density gradient and the molecular forms of AChE present in normal T lymphocytes and Jurkat cells.

Discussion

The expression of the ACHE gene in normal T lymphocytes and Jurkat E6-1 leukemic cells was analyzed using RT-PCR. AChE-H transcripts were detected in the first cell type, but AChE-T or AChE-R transcripts were not. These results agreed with a previous study of peripheral blood cells, including lymphocytes and erythrocytes, where the exclusive presence of the AChE-H transcript has been reported. This transcript encodes the AChE molecular forms of amphiphilic dimers and monomers (G2A and G1A) anchored to the plasma membrane by a glycosylphosphatidylinositol bond (17).

By contrast, the analysis of AChE transcripts in Jurkat leukemic cells using RT-PCR allowed the detection of AChE-H transcripts and the AChE-T variant, which demonstrated an alteration in the expression of the AChE gene in this cell line. This finding coincided with the results observed by Liao et al (58) regarding the expression of these transcripts in the Jurkat line.

In a previous study of P19 semidifferentiated neuronal cells via microarrays, overexpression of AChE-T transcripts was observed, which is associated with high expression levels of antiapoptotic genes and genes related to the splicing process (59). In Jurkat cells, the AChE-T transcript is probably involved in cell proliferation by regulating the expression of antiapoptotic agents and allowing the cell to survive.

The data obtained in the present study revealed that Jurkat cells presented a notable decrease in AChE activity compared with that of normal T lymphocytes. Other studies have reported lower AChE activity in cancer cells compared with normal cells (5,26,29). In the case of lung tumors, AChE activity decreases with increasing malignancy, which is related to an increase in ACh levels in these cells (25). The increase in AChE activity may be related to the development of hematological diseases and the regulation of immune function (35).

The function of lymphocytes is regulated not only by the cytokine system but also by the lymphoid cholinergic system independent of cholinergic neurons (3). A number of the components found in the nervous system, including ACh, choline acetyltransferase (ChAT), high-affinity choline transporter, muscarinic and nicotinic ACh receptors (mAChR and nAChR, respectively) and AChE, which together are known as the nonneuronal cholinergic system (NNCS), are expressed in lymphocytes (49,50,60). On the other hand, ACh and agonists of mAChRs and nAChRs have been reported to induce a variety of effects on lymphocytes, such as increases in the fluidity of the membrane, the synthesis of inositol triphosphate, the expression of the c-fos gene, the activation of DNA and RNA synthesis and cell proliferation. These findings support the hypothesis that ACh may regulate the functions of T lymphocytes (49,50,60).

A decrease in AChE activity can lead to an increase in the content of ACh present in the cell when it is not hydrolyzed. In the case of leukemic cells, the present study found that the specific activity of AChE was much lower than that of normal T lymphocytes. Similar results have been reported by Fujii et al (60), who documented that the levels of ACh in T lymphocytes are much lower than those in Jurkat cells. Thus, an increase in ACh levels in leukemic cells, produced by a lack of AChE activity, leads to an overregulation of the functions controlled by the NNCS, including the activation of cell proliferation (3). Fig. 7 shows a model that attempts to explain the induction of cell proliferation by ACh levels.

The sedimentation analysis of the molecular forms of AChE in T lymphocytes revealed the presence of a molecular structure corresponding to amphiphilic dimers and monomers (G2A and G1A), which are anchored to the membrane by GFI and originate through the expression of AChE-H transcripts (17,57,60). This finding is evidence of the expression of the AChE-H transcript detected in this type of cell. On the other hand, the sedimentation analysis of the AChE molecular forms in Jurkat cells revealed the presence of amphiphilic monomers and dimers (G2A and G1A) and hydrophilic tetramers (G4H). The molecular forms found in Jurkat cells were confirmed by the presence of the AChE-H transcripts responsible for the synthesis of the AChE G2A and G1A forms, as well as the expression of AChE-T transcripts from which the molecular form of AChE G4H is derived (57).

In this context, AChE activity has been proposed as a generic marker for small extracellular vesicles or exosomes. In Jurkat cells incubated with an extra depleted medium that can release AChE protein, a hydrophilic variant is released into the extracellular medium (58). AChE is an export protein whose synthesis is restarted in the rough endoplasmic reticulum, after which it is transported to the Golgi apparatus for modification and then secreted or deposited on the cell membrane (61).

Glycosylation variations were detected between both cell types, which suggested an alteration in the expression of AChE in a transformed cell compared with a normal cell. The primary sequence of the AChE protein contains three potential glycosylation sites linked to asparagine (Asn) residues 265, 350 and 464 through the glycosylation signal Asn-X-Ser (62). In mammals, these sites are conserved in all sequenced cholinesterases (63). When all three N-glycosylation signals of human AChE were suppressed by site-directed mutagenesis, all the sites were shown to be essential for effective biosynthesis and secretion. The extracellular AChE levels of mutants defective at one, two, or all three sites were 20-30%, 2-4% and ~0.5% of the wild-type level, respectively (62).

In a number of pathological states, the degree of AChE alteration includes posttranslational maturation, as evidenced by the modification of the glycosylation pattern of the AChE protein (25,29,64,65). A different glycosylation pattern of the AChE proteins contained in the S1+S2 extracts of normal T lymphocytes and Jurkat cells was observed when they interacted with lectins of different specificities. This finding indicated that the incorporation and/or removal of sugar residues during AChE maturation in the rough endoplasmic reticulum and the different lamellae of the Golgi apparatus were altered in the leukemic cell line, where a decrease in the incorporation of galactose and sialic acid into the AChE structure was mainly observed. AChE glycosylation is essential from the point of view of its activity and circulatory half-life (66). N-glycosylation may play a specific role in the unresolved noncholinergic roles of cholinesterases in cellular differentiation and development (67).

Sialic acid incorporation and its role in the pharmacokinetic properties of proteins was first reported by Morell et al (68), who discovered that sialylation alters the circulatory half-life of proteins by protecting the penultimate galactose residues, which are recognized by the asialoglycoprotein receptor (ASGPR). The binding of glycoproteins to ASGPR initiates protein degradation in the liver (69-71). Several studies have been built on the discovery by Morell et al (68) by increasing the level of sialic acid in various proteins, including cholinesterases, to prolong their circulatory retention time (62,71,72).

The present study was limited, using only one leukemia cell line (Jurkat). However, it was considered that the results obtained are a valuable starting point for future studies. The authors continue to investigate the possible participation of AChE and other components of the nonneuronal cholinergic system in cell growth and proliferation, allowing the development of new proposals for therapeutic targets against acute lymphoblastic leukemia.

Future research will involve a wider range of leukemia cell lines and clinical samples. Additionally, it will include data on activated normal T lymphocytes to gain a comprehensive understanding of the role of AChE in different functional states of lymphocytes. Notably, future studies of AChE expression in samples from patients with leukemia may allow the use of AChE as a tool for the early detection of leukemia, with the aim on increasing the effectiveness of current treatments.

The results of the present study showed that the content and composition of AChE are altered in Jurkat leukemic cells compared with normal T lymphocytes, from the change in the expression pattern of the transcripts to the decrease in the specific activity of AChE, the differences in glycan processing and the alteration in the assembly of molecular shapes. This opens the way to explore the possible participation of AChE in the proliferation and differentiation of lymphoid cells, as well as in the development of new therapeutic targets against acute lymphoblastic leukemia and the creation of timely detection methods. Further studies should be conducted in other leukemia cell types to elucidate the underlying mechanisms involved in the generation of the proposed molecular targets.

Acknowledgements

The authors thank Dr María Concepción Gutiérrez Ruiz, Department of Health Sciences, Metropolitan Autonomous University-Iztapalapa Campus for providing the RNA samples from the HepG2 cell line; Dr Rocio Salceda Sacanelles, National Autonomous University of Mexico and Dr Hector Serrano, Metropolitan Autonomous University-Iztapalapa Campus for their technical assistance.

Funding

Funding: The present study was funded by the E022 Program: Recursos Fiscales para Investigación, Instituto Nacional de Pediatría (grant no. 2022/067 LF-L and 2019/072 SE-F). It was also supported by Consejo Nacional de Humanidades, Ciencias y Tecnologías, México (UAM-I: grant nos. 309-0; C/PFPN-2002-35-32) and Ciencia de Frontera 2023 (grant no. CONAHCYT CF-2023-I-2755 and CF-2023-I-811).

Availability of data and materials

The data generated in the present study may be requested from the corresponding author.

Authors' contributions

Data acquisition, processing and validation was performed by LF-L, JG-O, SE-F, GL-V, ID-D, RV-R and IG-T; the study investigation was carried out by LF-L, JG-O and SE-F carried out project administration; LF-L and JG-O wrote the original draft of the manuscript; LF-L, JG-O, SE-F, GL-V, ID-D and IG-T reviewed and edited the manuscript. SE-F, GL-V, ID-D, IG-T, RL-D, RV-R, RV-Q and MM confirm the authenticity of all the raw data. All authors read and approved the final version of the manuscript.

Ethics approval and consent to participate

The research ethics committee of the National Institute of Pediatrics approved the use of blood samples donated by healthy adults with a registration in the National Bioethics Commission with the number CONBIOETICA-09-CEI-025-20161215 of the National Institute of Pediatrics, Mexico.

Patient consent for publication

Not applicable.

Competing interests

The authors declare that they have no competing interests.

References

1 

Soreq H and Seidman S: Acetylcholinesterase-new roles for an old actor. Nat Rev Neurosci. 2:294–302. 2001.PubMed/NCBI View Article : Google Scholar

2 

Grisaru D, Pick M, Perry C, Sklan EH, Almog R, Goldberg I, Naparstek E, Lessing JB, Soreq H and Deutsch V: Hydrolytic and nonenzymatic functions of acetylcholinesterase comodulate hemopoietic stress responses. J Immunol. 176:27–35. 2006.PubMed/NCBI View Article : Google Scholar

3 

Fujii T, Mashimo M, Moriwaki Y, Misawa H, Ono S, Horiguchi K and Kawashima K: Expression and function of the cholinergic system in immune cells. Front Immunol. 8(1085)2017.PubMed/NCBI View Article : Google Scholar

4 

Onganer PU, Djamgoz MBA, Whyte K and Greenfield SA: An acetylcholinesterase-derived peptide inhibits endocytic membrane activity in a human metastatic breast cancer cell line. Biochim Biophys Acta. 1760:415–420. 2006.PubMed/NCBI View Article : Google Scholar

5 

Pérez-Aguilar B, Vidal CJ, Palomec G, García-Dolores F, Gutiérrez-Ruiz MC, Bucio L, Gómez-Olivares JL and Gómez-Quiroz LE: Acetylcholinesterase is associated with a decrease in cell proliferation of hepatocellular carcinoma cells. Biochim Biophys Acta. 1852:1380–1387. 2015.PubMed/NCBI View Article : Google Scholar

6 

Villeda-González JD, Gómez-Olivares JL, Baiza-Gutman LA, Manuel-Apolinar L, Damasio-Santana L, Millán-Pacheco C, Ángeles-Mejía S, Cortés-Ginez MC, Cruz-López M, Vidal-Moreno CJ and Díaz-Flores M: Nicotinamide reduces inflammation and oxidative stress via the cholinergic system in fructose-induced metabolic syndrome in rats. Life Sci. 250(117585)2020.PubMed/NCBI View Article : Google Scholar

7 

Zhang JY, Jiang H, Gao W, Wu J, Peng K, Shi YF and Zhang XJ: The JNK/AP1/ATF2 pathway is involved in H2O2-induced acetylcholinesterase expression during apoptosis. Cell Mol Life Sci. 65:1435–1445. 2008.PubMed/NCBI View Article : Google Scholar

8 

Zhang XJ, Yang L, Zhao Q, Caen JP, He HY, Jin QH, Guo LH, Alemany M, Zhang LY and Shi YF: Induction of acetylcholinesterase expression during apoptosis in various cell types. Cell Death Differ. 9:790–800. 2002.PubMed/NCBI View Article : Google Scholar

9 

Getman DK, Eubanks JH, Camp S, Evans GA and Taylor P: The human gene encoding acetylcholinesterase is located on the long arm of chromosome 7. Am J Hum Genet. 51:170–177. 1992.PubMed/NCBI

10 

Taylor P and Radić Z: The cholinesterases: From genes to proteins. Annu Rev Pharmacol Toxicol. 34:281–320. 1994.PubMed/NCBI View Article : Google Scholar

11 

Soreq H, Ben-Aziz R, Prody CA, Seidman S, Gnatt A, Neville L, Lieman-Hurwitz J, Lev-Lehman E, Ginzberg D, Lipidot-Lifson Y, et al: Molecular cloning and construction of the coding region for human acetylcholinesterase reveals a G + C-rich attenuating structure. Proc Natl Acad Sci USA. 87:9688–9692. 1990.PubMed/NCBI View Article : Google Scholar

12 

Massoulié J, Pezzementi L, Bon S, Krejci E and Vallette FM: Molecular and cellular biology of cholinesterases. Prog Neurobiol. 41:31–91. 1993.PubMed/NCBI View Article : Google Scholar

13 

Massoulié J, Anselmet A, Bon S, Krejci E, Legay C, Morel N and Simon S: Acetylcholinesterase: C-terminal domains, molecular forms and functional localization. J Physiol Paris. 92:183–190. 1998.PubMed/NCBI View Article : Google Scholar

14 

Dori A and Soreq H: ARP, the cleavable C-terminal peptide of ‘readthrough’ acetylcholinesterase, promotes neuronal development and plasticity. J Mol Neurosci. 28:247–255. 2006.PubMed/NCBI View Article : Google Scholar

15 

Perrier NA, Salani M, Falasca C, Bon S, Augusti-Tocco G and Massoulié J: The readthrough variant of acetylcholinesterase remains very minor after heat shock, organophosphate inhibition and stress, in cell culture and in vivo. J Neurochem. 94:629–638. 2005.PubMed/NCBI View Article : Google Scholar

16 

Deutsch VR, Pick M, Perry C, Grisaru D, Hemo Y, Golan-Hadari D, Grant A, Eldor A and Soreq H: The stress-associated acetylcholinesterase variant AChE-R is expressed in human CD34(+) hematopoietic progenitors and its C-terminal peptide ARP promotes their proliferation. Exp Hematol. 30:1153–1161. 2002.PubMed/NCBI View Article : Google Scholar

17 

Pick M, Flores-Flores C, Grisaru D, Shochat S, Deutsch V and Soreq H: Blood-cell-specific acetylcholinesterase splice variations under changing stimuli. Int J Dev Neurosci. 22:523–531. 2004.PubMed/NCBI View Article : Google Scholar

18 

Grisaru D, Sternfeld M, Eldor A, Glick D and Soreq H: Structural roles of acetylcholinesterase variants in biology and pathology. Eur J Biochem. 264:672–686. 1999.PubMed/NCBI View Article : Google Scholar

19 

Sternfeld M, Shoham S, Klein O, Flores-Flores C, Evron T, Idelson GH, Kitsberg D, Patrick JW and Soreq H: Excess ‘read-through’ acetylcholinesterase attenuates but the ‘synaptic’ variant intensifies neurodeterioration correlates. Proc Natl Acad Sci USA. 97:8647–8652. 2000.PubMed/NCBI View Article : Google Scholar

20 

Fischer K, Brown J, Scherer SW, Schramm P, Stewart J, Fugazza G, Pascheberg U, Peter W, Tsui LC, Lichter P and Döhner H: Delineation of genomic regions in chromosome band 7q22 commonly deleted in myeloid leukemias. Recent Results Cancer Res. 144:46–52. 1998.PubMed/NCBI View Article : Google Scholar

21 

Zeng WR, Watson P, Lin J, Jothy S, Lidereau R, Park M and Nepveu A: Refined mapping of the region of loss of heterozygosity on the long arm of chromosome 7 in human breast cancer defines the location of a second tumor suppressor gene at 7q22 in the region of the CUTL1 gene. Oncogene. 18:2015–2021. 1999.PubMed/NCBI View Article : Google Scholar

22 

Neville PJ, Thomas N and Campbell IG: Loss of heterozygosity at 7q22 and mutation analysis of the CDP gene in human epithelial ovarian tumors. Int J Cancer. 91:345–349. 2001.PubMed/NCBI View Article : Google Scholar

23 

Sáez-Valero J and Vidal CJ: Biochemical properties of acetyl- and butyrylcholinesterase in human meningioma. Biochim Biophys Acta. 1317:210–218. 1996.PubMed/NCBI View Article : Google Scholar

24 

Vidal CJ: Expression of cholinesterases in brain and non-brain tumours. Chem Biol Interact. 157-158:227–232. 2005.PubMed/NCBI View Article : Google Scholar

25 

Martínez-Moreno P, Nieto-Cerón S, Torres-Lanzas J, Ruiz-Espejo F, Tovar-Zapata I, Martínez-Hernández P, Rodríguez-López JN, Vidal CJ and Cabezas-Herrera J: Cholinesterase activity of human lung tumours varies according to their histological classification. Carcinogenesis. 27:429–436. 2006.PubMed/NCBI View Article : Google Scholar

26 

Montenegro MF, Nieto-Cerón S, Ruiz-Espejo F, Páez de la Cadena M, Rodríguez-Berrocal FJ and Vidal CJ: Cholinesterase activity and enzyme components in healthy and cancerous human colorectal sections. Chem Biol Interact. 157-158:429–430. 2005.PubMed/NCBI View Article : Google Scholar

27 

Montenegro MF, Ruiz-Espejo F, Campoy FJ, Muñoz-Delgado E, de la Cadena MP, Rodríguez-Berrocal FJ and Vidal CJ: Cholinesterases are down-expressed in human colorectal carcinoma. Cell Mol Life Sci. 63:2175–2182. 2006.PubMed/NCBI View Article : Google Scholar

28 

Martínez-López de Castro A, Nieto-Cerón S, Aurelio PC, Galbis-Martínez L, Latour-Pérez J, Torres-Lanzas J, Tovar-Zapata I, Martínez-Hernández P, RodríRodríguez-López JN and Cabezas-Herrera J: Cancer-associated differences in acetylcholinesterase activity in bronchial aspirates from patients with lung cancer. Clin Sci (Lond). 115:245–253. 2008.PubMed/NCBI View Article : Google Scholar

29 

Ruiz-Espejo F, Cabezas-Herrera J, Illana J, Campoy FJ, Muñoz-Delgado E and Vidal CJ: Breast cancer metastasis alters acetylcholinesterase activity and the composition of enzyme forms in axillary lymph nodes. Breast Cancer Res Treat. 80:105–114. 2003.PubMed/NCBI View Article : Google Scholar

30 

García-Ayllón MS, Sáez-Valero J, Muñoz-Delgado E and Vidal CJ: Identification of hybrid cholinesterase forms consisting of acetyl- and butyrylcholinesterase subunits in human glioma. Neuroscience. 107:199–208. 2001.PubMed/NCBI View Article : Google Scholar

31 

Zhao Y, Wang X, Wang T, Hu X, Hui X, Yan M, Gao Q, Chen T, Li J, Yao M, et al: Acetylcholinesterase, a key prognostic predictor for hepatocellular carcinoma, suppresses cell growth and induces chemosensitization. Hepatology. 53:493–503. 2011.PubMed/NCBI View Article : Google Scholar

32 

Xu H, Shen Z, Xiao J, Yang Y, Huang W, Zhou Z, Shen J, Zhu Y, Liu XY and Chu L: Acetylcholinesterase overexpression mediated by oncolytic adenovirus exhibited potent anti-tumor effect. BMC Cancer. 14(668)2014.PubMed/NCBI View Article : Google Scholar

33 

Castillo-González AC, Nieto-Cerón S, Pelegrín-Hernández JP, Montenegro MF, Noguera JA, López-Moreno MF, Rodríguez-López JN, Vidal CJ, Hellín-Meseguer D and Cabezas-Herrera J: Dysregulated cholinergic network as a novel biomarker of poor prognostic in patients with head and neck squamous cell carcinoma. BMC Cancer. 15(385)2015.PubMed/NCBI View Article : Google Scholar

34 

Castillo-González AC, Pelegrín-Hernández JP, Nieto-Cerón S, Madrona AP, Noguera JA, López-Moreno MF, Rodríguez-López JN, Vidal CJ, Hellín-Meseguer D and Cabezas-Herrera J: Unbalanced acetylcholinesterase activity in larynx squamous cell carcinoma. Int Immunopharmacol. 29:81–86. 2015.PubMed/NCBI View Article : Google Scholar

35 

Battisti V, Bagatini MD, Maders LDK, Chiesa J, Santos KF, Gonçalves JF, Abdalla FH, Battisti IE, Schetinger MR and Morsch VM: Cholinesterase activities and biochemical determinations in patients with prostate cancer: Influence of Gleason score, treatment and bone metastasis. Biomed Pharmacother. 66:249–255. 2012.PubMed/NCBI View Article : Google Scholar

36 

Yu H, Xia H, Tang Q, Xu H, Wei G, Chen Y, Dai X, Gong Q and Bi F: Acetylcholine acts through M3 muscarinic receptor to activate the EGFR signaling and promotes gastric cancer cell proliferation. Sci Rep. 7(40802)2017.PubMed/NCBI View Article : Google Scholar

37 

Resende RR, Alves AS, Britto LRG and Ulrich H: Role of acetylcholine receptors in proliferation and differentiation of P19 embryonal carcinoma cells. Exp Cell Res. 314:1429–1443. 2008.PubMed/NCBI View Article : Google Scholar

38 

Cheng K, Samimi R, Xie G, Shant J, Drachenberg C, Wade M, Davis RJ, Nomikos G and Raufman JP: Acetylcholine release by human colon cancer cells mediates autocrine stimulation of cell proliferation. Am J Physiol Gastrointest Liver Physiol. 295:G591–G597. 2008.PubMed/NCBI View Article : Google Scholar

39 

Pettersson A, Nilsson L, Nylund G, Khorram-Manesh A, Nordgren S and Delbro DS: Is acetylcholine an autocrine/paracrine growth factor via the nicotinic alpha7-receptor subtype in the human colon cancer cell line HT-29? Eur J Pharmacol. 609:27–33. 2009.PubMed/NCBI View Article : Google Scholar

40 

Song P, Sekhon HS, Jia Y, Keller JA, Blusztajn JK, Mark GP and Spindel ER: Acetylcholine is synthesized by and acts as an autocrine growth factor for small cell lung carcinoma. Cancer Res. 63:214–221. 2003.PubMed/NCBI

41 

Lev-Lehman E, Ginzberg D, Hornreich G, Ehrlich G, Meshorer A, Eckstein F, Soreq H and Zakut H: Antisense inhibition of acetylcholinesterase gene expression causes transient hematopoietic alterations in vivo. Gene Ther. 1:127–135. 1994.PubMed/NCBI

42 

Soreq H, Patinkin D, Lev-Lehman E, Grifman M, Ginzberg D, Eckstein F and Zakut H: Antisense oligonucleotide inhibition of acetylcholinesterase gene expression induces progenitor cell expansion and suppresses hematopoietic apoptosis ex vivo. Proc Natl Acad Sci USA. 91:7907–7911. 1994.PubMed/NCBI View Article : Google Scholar

43 

Brown LM, Blair A, Gibson R, Everett GD, Cantor KP, Schuman LM, Burmeister LF, Van Lier SF and Dick F: Pesticide exposures and other agricultural risk factors for leukemia among men in Iowa and Minnesota. Cancer Res. 50:6585–6591. 1990.PubMed/NCBI

44 

Cantor KP, Blair A, Everett G, Gibson R, Burmeister LF, Brown LM, Schuman L and Dick FR: Pesticides and other agricultural risk factors for non-Hodgkin's lymphoma among men in Iowa and Minnesota. Cancer Res. 52:2447–2455. 1992.PubMed/NCBI

45 

Cabello G, Valenzuela M, Vilaxa A, Durán V, Rudolph I, Hrepic N and Calaf G: A rat mammary tumor model induced by the organophosphorous pesticides parathion and malathion, possibly through acetylcholinesterase inhibition. Environ Health Perspect. 109:471–479. 2001.PubMed/NCBI View Article : Google Scholar

46 

Nasterlack M: Pesticides and childhood cancer: An update. Int J Hyg Environ Health. 210:645–657. 2007.PubMed/NCBI View Article : Google Scholar

47 

Calaf GM, Parra E and Garrido F: Cell proliferation and tumor formation induced by eserine, an acetylcholinesterase inhibitor, in rat mammary gland. Oncol Rep. 17:25–33. 2007.PubMed/NCBI View Article : Google Scholar

48 

Kawashima K and Fujii T: Extraneuronal cholinergic system in lymphocytes. Pharmacol Ther. 86:29–48. 2000.PubMed/NCBI View Article : Google Scholar

49 

Kawashima K and Fujii T: The lymphocytic cholinergic system and its biological function. Life Sci. 72:2101–2109. 2003.PubMed/NCBI View Article : Google Scholar

50 

Kawashima K and Fujii T: Expression of non-neuronal acetylcholine in lymphocytes and its contribution to the regulation of immune function. Front Biosci. 9:2063–2085. 2004.PubMed/NCBI View Article : Google Scholar

51 

Ellman GL, Courtney KD, Andres V Jr and Feather-Stone RM: A new and rapid colorimetric determination of acetylcholinesterase activity. Biochem Pharmacol. 7:88–95. 1961.PubMed/NCBI View Article : Google Scholar

52 

Bradford MM: A rapid and sensitive method for the quantitation of microgram quantities of protein utilizing the principle of protein-dye binding. Anal Biochem. 72:248–254. 1976.PubMed/NCBI View Article : Google Scholar

53 

Martin RG and Ames BN: A method for determining the sedimentation behavior of enzymes: Application to protein mixtures. J Biol Chem. 236:1372–1379. 1961.PubMed/NCBI

54 

Bartha E, Rakonczay Z, Kása P, Hollán S and Gyévai A: Molecular form of human lymphocyte membrane-bound acetylcholinesterase. Life Sci. 41:1853–1860. 1987.PubMed/NCBI View Article : Google Scholar

55 

Sine JP and Caye-Vaugien C: Properties and characterization of soluble forms of lymphocyte acetylcholinesterase from an ox. Biochimie. 66:203–214. 1984.PubMed/NCBI View Article : Google Scholar : (In French).

56 

Toutant JP, Richards MK, Krall JA and Rosenberry TL: Molecular forms of acetylcholinesterase in two sublines of human erythroleukemia K562 cells. Sensitivity or resistance to phosphatidylinositol-specific phospholipase C and biosynthesis. Eur J Biochem. 187:31–38. 1990.PubMed/NCBI View Article : Google Scholar

57 

Massoulié J, Bon S, Perrier N and Falasca C: The C-terminal peptides of acetylcholinesterase: Cellular trafficking, oligomerization and functional anchoring. Chem Biol Interact. 157-158:3–14. 2005.PubMed/NCBI View Article : Google Scholar

58 

Liao Z, Jaular LM, Soueidi E, Jouve M, Muth DC, Schøyen TH, Seale T, Haughey NJ, Ostrowski M, Théry C and Witwer KW: Acetylcholinesterase is not a generic marker of extracellular vesicles. J Extracell Vesicles. 8(1628592)2019.PubMed/NCBI View Article : Google Scholar

59 

Ben-Ari S, Toiber D, Sas AS, Soreq H and Ben-Shaul Y: Modulated splicing-associated gene expression in P19 cells expressing distinct acetylcholinesterase splice variants. J Neurochem. 97 (Suppl 1):S24–S34. 2006.PubMed/NCBI View Article : Google Scholar

60 

Fujii T, Takada-Takatori Y and Kawashima K: Basic and clinical aspects of non-neuronal acetylcholine: Expression of an independent, non-neuronal cholinergic system in lymphocytes and its clinical significance in immunotherapy. J Pharmacol Sci. 106:186–192. 2008.PubMed/NCBI View Article : Google Scholar

61 

Jin QH, He HY, Shi YF, Lu H and Zhang XJ: Overexpression of acetylcholinesterase inhibited cell proliferation and promoted apoptosis in NRK cells. Acta Pharmacol Sin. 25:1013–1021. 2004.PubMed/NCBI

62 

Velan B, Kronman C, Ordentlich A, Flashner Y, Leitner M, Cohen S and Shafferman A: N-glycosylation of human acetylcholinesterase: Effects on activity, stability and biosynthesis. Biochem J. 296:649–656. 1993.PubMed/NCBI View Article : Google Scholar

63 

Doctor BP, Chapman TC, Christner CE, Deal CD, De La Hoz DM, Gentry MK, Ogert RA, Rush RS, Smyth KK and Wolfe AD: Complete amino acid sequence of fetal bovine serum acetylcholinesterase and its comparison in various regions with other cholinesterases. FEBS Lett. 266:123–127. 1990.PubMed/NCBI View Article : Google Scholar

64 

Campoy FJ, Cabezas-Herrera J and Vidal CJ: Interaction of AChE with Lens culinaris agglutinin reveals differences in glycosylation of molecular forms in sarcoplasmic reticulum membrane subfractions. J Neurosci Res. 33:568–578. 1992.PubMed/NCBI View Article : Google Scholar

65 

Cabezas-Herrera J, Moral-Naranjo MT, Campoy FJ and Vidal CJ: G4 forms of acetylcholinesterase and butyrylcholinesterase in normal and dystrophic mouse muscle differ in their interaction with Ricinus communis agglutinin. Biochim Biophys Acta. 1225:283–288. 1994.PubMed/NCBI View Article : Google Scholar

66 

Nalivaeva NN and Turner AJ: Post-translational modifications of proteins: Acetylcholinesterase as a model system. Proteomics. 1:735–747. 2001.PubMed/NCBI View Article : Google Scholar

67 

Patinkin D, Seidman S, Eckstein F, Benseler F, Zakut H and Soreq H: Manipulations of cholinesterase gene expression modulate murine megakaryocytopoiesis in vitro. Mol Cell Biol. 10:6046–6050. 1990.PubMed/NCBI View Article : Google Scholar

68 

Morell AG, Irvine RA, Sternlieb I, Scheinberg IH and Ashwell G: Physical and chemical studies on ceruloplasmin. V. Metabolic studies on sialic acid-free ceruloplasmin in vivo. J Biol Chem. 243:155–159. 1968.PubMed/NCBI

69 

Ashwell G and Harford J: Carbohydrate-specific receptors of the liver. Annu Rev Biochem. 51:531–554. 1982.PubMed/NCBI View Article : Google Scholar

70 

Bon C, Hofer T, Bousquet-Mélou A, Davies MR and Krippendorff BF: Capacity limits of asialoglycoprotein receptor-mediated liver targeting. MAbs. 9:1360–1369. 2017.PubMed/NCBI View Article : Google Scholar

71 

Chia S, Tay SJ, Song Z, Yang Y, Walsh I and Pang KT: Enhancing pharmacokinetic and pharmacodynamic properties of recombinant therapeutic proteins by manipulation of sialic acid content. Biomed Pharmacother. 163(114757)2023.PubMed/NCBI View Article : Google Scholar

72 

Kronman C, Velan B, Marcus D, Ordentlich A, Reuveny S and Shafferman A: Involvement of oligomerization, N-glycosylation and sialylation in the clearance of cholinesterases from the circulation. Biochem J. 311:959–967. 1995.PubMed/NCBI View Article : Google Scholar

Related Articles

Journal Cover

November-2024
Volume 21 Issue 5

Print ISSN: 2049-9434
Online ISSN:2049-9442

Sign up for eToc alerts

Recommend to Library

Copy and paste a formatted citation
x
Spandidos Publications style
Gómez‑Olivares JL, López‑Durán RM, Enríquez‑Flores S, López‑Velázquez G, De La Mora‑De La Mora I, García‑Torres I, Viedma‑Rodríguez R, Valencia‑Quintana R, Milić M, Flores‑López LA, Flores‑López LA, et al: Preliminary exploration of the expression of acetylcholinesterase in normal human T lymphocytes and leukemic Jurkat T cells. Biomed Rep 21: 158, 2024.
APA
Gómez‑Olivares, J.L., López‑Durán, R.M., Enríquez‑Flores, S., López‑Velázquez, G., De La Mora‑De La Mora, I., García‑Torres, I. ... Flores‑López, L.A. (2024). Preliminary exploration of the expression of acetylcholinesterase in normal human T lymphocytes and leukemic Jurkat T cells. Biomedical Reports, 21, 158. https://doi.org/10.3892/br.2024.1846
MLA
Gómez‑Olivares, J. L., López‑Durán, R. M., Enríquez‑Flores, S., López‑Velázquez, G., De La Mora‑De La Mora, I., García‑Torres, I., Viedma‑Rodríguez, R., Valencia‑Quintana, R., Milić, M., Flores‑López, L. A."Preliminary exploration of the expression of acetylcholinesterase in normal human T lymphocytes and leukemic Jurkat T cells". Biomedical Reports 21.5 (2024): 158.
Chicago
Gómez‑Olivares, J. L., López‑Durán, R. M., Enríquez‑Flores, S., López‑Velázquez, G., De La Mora‑De La Mora, I., García‑Torres, I., Viedma‑Rodríguez, R., Valencia‑Quintana, R., Milić, M., Flores‑López, L. A."Preliminary exploration of the expression of acetylcholinesterase in normal human T lymphocytes and leukemic Jurkat T cells". Biomedical Reports 21, no. 5 (2024): 158. https://doi.org/10.3892/br.2024.1846