Skip to main content

Methods for the Measurement of Early Events in Toxoplasma gondii Immunity in Mouse Cells

  • Protocol
  • First Online:
Toxoplasma gondii

Abstract

Critical steps in resistance of mice against Toxoplasma gondii occur in the first 2 or 3 h after the pathogen has entered a cell that has been exposed to interferon γ (IFNγ). The newly formed parasitophorous vacuole is attacked by the IFNγ-inducible IRG proteins and disrupted, resulting in death of the parasite and necrotic death of the cell. Here we describe some techniques that we have used to describe and quantify these events in different combinations of the host and the parasite.

This is a preview of subscription content, log in via an institution to check access.

Access this chapter

Protocol
USD 49.95
Price excludes VAT (USA)
  • Available as PDF
  • Read on any device
  • Instant download
  • Own it forever
eBook
USD 89.00
Price excludes VAT (USA)
  • Available as EPUB and PDF
  • Read on any device
  • Instant download
  • Own it forever
Softcover Book
USD 119.99
Price excludes VAT (USA)
  • Compact, lightweight edition
  • Dispatched in 3 to 5 business days
  • Free shipping worldwide - see info
Hardcover Book
USD 169.99
Price excludes VAT (USA)
  • Durable hardcover edition
  • Dispatched in 3 to 5 business days
  • Free shipping worldwide - see info

Tax calculation will be finalised at checkout

Purchases are for personal use only

Institutional subscriptions

Abbreviations

CO2:

Carbon dioxide

DMEM:

Dulbecco’s modified Eagle’s medium

EDTA:

Ethylenediaminetetraacetic acid

FACS:

Fluorescence-activated cell sorting

FBS:

Fetal bovine serum

H2O2:

Hydrogen peroxide

HEPES:

4-(2-Hydroxyethyl)-1-piperazineethanesulfonic acid)

HFFs:

Human foreskin fibroblasts = Hs27 cells

HI-FBS:

Heat-inactivated fetal bovine serum

IFNγ:

Interferon (IFN)-gamma

MMS:

Multichannel microscope slide

MOI:

Multiplicity of infection

NEAA:

Nonessential amino acids

P/W buffer:

Permeabilization/wash buffer (for intracellular staining for FACS)

PBS:

Phosphate-buffered saline

PFA:

Paraformaldehyde

RCF:

Relative centrifugal force

RT:

Room temperature

SAG1:

Toxoplasma major surface antigen 1 or P30

Sta:

Staurosporine

T. gondii, Tg:

Toxoplasma gondii

WD:

Working dilution

References

  1. Müller UB, Howard JC (2016) The impact of Toxoplasma gondii on the mammalian genome. Curr Opin Microbiol 32:19–25

    Article  PubMed  CAS  Google Scholar 

  2. Carruthers V, Boothroyd JC (2007) Pulling together: an integrated model of Toxoplasma cell invasion. Curr Opin Microbiol 10:83–89

    Article  CAS  PubMed  Google Scholar 

  3. Besteiro S, Dubremetz JF, Lebrun M (2011) The moving junction of apicomplexan parasites: a key structure for invasion. Cell Microbiol 13:797–805

    Article  CAS  PubMed  Google Scholar 

  4. Dubremetz JF (2007) Rhoptries are major players in Toxoplasma gondii invasion and host cell interaction. Cell Microbiol 9:841–848

    Article  CAS  PubMed  Google Scholar 

  5. Boothroyd JC, Dubremetz J-F (2008) Kiss and spit: the dual roles of Toxoplasma rhoptries. Nat Rev Microbiol 6:79

    Article  CAS  PubMed  Google Scholar 

  6. Martens S et al (2005) Disruption of Toxoplasma gondii parasitophorous vacuoles by the mouse p47-resistance GTPases. PLoS Pathog 1:0187–0201

    Article  CAS  Google Scholar 

  7. Zhao YO, Khaminets A, Hunn JP, Howard JC (2009) Disruption of the Toxoplasma gondii parasitophorous vacuole by IFNg-inducible immunity-related GTPases (IRG proteins) triggers necrotic cell death. PLoS Pathog 5:e1000288

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  8. Fentress SJ et al (2011) Phosphorylation of immunity-related GTPases by a Toxoplasma gondii secreted kinase promotes macrophage survival and virulence. Cell Host Microbe 8:484–495

    Article  CAS  Google Scholar 

  9. Khaminets A et al (2010) Coordinated loading of IRG resistance GTPases on to the Toxoplasma gondii parasitophorous vacuole. Cell Microbiol 12:939–961

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  10. Weiss LM, Kim K (2013) Toxoplasma gondii: the model apicomplexan – perspectives and methods: second edition. Elsevier, Amsterdam

    Google Scholar 

  11. Allen I (2013) Mouse models of innate immunity: methods and protocols, vol 1031. Humana Press, New York, NY

    Book  Google Scholar 

  12. Antony VB, Owen CL, Hadley KJ (1989) Pleural mesothelial cells stimulated by asbestos release chemotactic activity for neutrophils in vitro. Am Rev Respir Dis 139:199–206

    Article  CAS  PubMed  Google Scholar 

  13. Lilue J, Müller UB, Steinfeldt T, Howard JC (2013) Reciprocal virulence and resistance polymorphism in the relationship between Toxoplasma gondii and the house mouse. elife 2013:1–21

    Google Scholar 

  14. Incella (2016) ScreenFect®A transfection reagent protocol. Incella, Baden-Württemberg, pp 1–5

    Google Scholar 

  15. Southern PJ, Berg P (1982) Transformation of mammalian cells to antibiotic resistance with a bacterial gene under control of the SV40 early region promoter. J Mol Appl Genet 1(4):327–341

    CAS  PubMed  Google Scholar 

  16. Chang PL et al (1986) Transformation of human cultured fibroblasts with plasmids carrying dominant selection markers and immortalizing potential. Exp Cell Res 167:407–416

    Article  CAS  PubMed  Google Scholar 

  17. Li LM et al (2015) ScreenFect A: an efficient and low toxic liposome for gene delivery to mesenchymal stem cells. Int J Pharm 488:1–11

    Article  CAS  PubMed  Google Scholar 

  18. Li L et al (2012) A biomimetic lipid library for gene delivery through thiol-yne click chemistry. Biomaterials 33:8160–8166

    Article  CAS  PubMed  Google Scholar 

  19. Fisch DH et al (2018) An artificial intelligence workflow for defining host-pathogen interactions. bioRxiv 408450. https://doi.org/10.1101/408450

  20. Schindelin J et al (2012) Fiji – an open source platform for biological image analysis. Nat Methods 9:676–682

    Article  CAS  PubMed  Google Scholar 

  21. Papic N, Hunn JP, Pawlowski N, Zerrahn J, Howard JC (2008) Inactive and active states of the interferon-inducible resistance GTPase, Irga6, in vivo. J Biol Chem 283:32143–32151

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  22. Lotze MT, Tracey KJ (2005) High-mobility group box 1 protein (HMGB1): nuclear weapon in the immune arsenal. Nat Rev Immunol 5:331

    Article  CAS  PubMed  Google Scholar 

  23. Zhao YO et al (2009) Toxoplasma gondii and the immunity-related GTPase (IRG) resistance system in mice – a review. Mem Inst Oswaldo Cruz 104:234–240

    Article  CAS  PubMed  Google Scholar 

  24. Scaffidi P, Misteli T, Bianchi ME (2002) Release of chromatin protein HMGB1 by necrotic cells triggers inflammation. Nature 418:191–195

    Article  CAS  PubMed  Google Scholar 

  25. Bonaldi T et al (2003) Monocytic cells hyperacetylate chromatin protein HMGB1 to redirect it towards secretion. EMBO 22:5551–5560

    Article  CAS  Google Scholar 

  26. Andersson U et al (2000) High mobility group 1 protein (Hmg-1) stimulates proinflammatory cytokine synthesis in human monocytes. J Exp Med 192:565–570

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  27. Tang D et al (2010) Endogenous HMGB1 regulates autophagy. J Cell Biol 190:881–892

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  28. Vanden Berghe T et al (2010) Necroptosis, necrosis and secondary necrosis converge on similar cellular disintegration features. Cell Death Differ 17:922–930

    Article  CAS  Google Scholar 

  29. Bertho ÁL, Santiago MA, Coutinho SG (2000) Flow cytometry in the study of cell death. Mem Inst Oswaldo Cruz 95:429–433

    Article  CAS  PubMed  Google Scholar 

  30. Holmes KL et al (2014) International society for the advancement of cytometry cell sorter biosafety standards. Citometry A 85:434–453

    Article  Google Scholar 

  31. Könen-Waisman S, Howard JC (2007) Cell-autonomous immunity to Toxoplasma gondii in mouse and man. Microbes Infect 9:1652–1661

    Article  PubMed  CAS  Google Scholar 

  32. Martens S et al (2004) Mechanisms regulating the positioning of mouse p47 resistance GTPases LRG-47 and IIGP1 on cellular membranes: retargeting to plasma membrane induced by phagocytosis. J Immunol 173:2594–2606

    Article  CAS  PubMed  Google Scholar 

  33. Zerrahn J, Schaible UE, Brinkmann V, Guhlich U, Kaufmann SHE (2002) The IFN-inducible Golgi- and endoplasmic reticulum-associated 47-kDa GTPase IIGP is transiently expressed during listeriosis. J Immunol 168:3428–3436

    Article  CAS  PubMed  Google Scholar 

  34. Carlow DA et al (1998) Specific antiviral activity demonstrated by TGTP, a member of a new family of interferon-induced GTPases. J Immunol 161:2348–2355

    CAS  PubMed  Google Scholar 

  35. Maric-Biresev J et al (2016) Loss of the interferon-γ-inducible regulatory immunity-related GTPase (IRG), Irgm1, causes activation of effector IRG proteins on lysosomes, damaging lysosomal function and predicting the dramatic susceptibility of Irgm1-deficient mice to infection. BMC Biol 14:1–20

    Article  CAS  Google Scholar 

  36. Pawlowski N et al (2011) The activation mechanism of Irga6, an interferon-inducible GTPase contributing to mouse resistance against Toxoplasma gondii. BMC Biol 9:7

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  37. Steinfeldt T et al (2010) Phosphorylation of mouse immunity-related gtpase (IRG) resistance proteins is an evasion strategy for virulent Toxoplasma gondii. PLoS Biol 8:e1000576

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  38. Butcher BA et al (2005) p47 GTPases regulate Toxoplasma gondii survival in sctivated macrophages. Infect Immun 73:3278–3286

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  39. Springer HM, Schramm M, Taylor G a, Howard JC (2013) Irgm1 (LRG-47), a regulator of cell-autonomous immunity, does not localize to mycobacterial or listerial phagosomes in IFN-γ-induced mouse cells. J Immunol 191:1765–1774

    Article  CAS  PubMed  Google Scholar 

  40. Fleckenstein MC et al (2012) A Toxoplasma gondii pseudokinase inhibits host irg resistance proteins. PLoS Biol 10:14

    Article  CAS  Google Scholar 

  41. Naviaux RK, Costanzi E, Haas M, Verma IM (1996) The pCL vector system: rapid production of helper-free, high-titer, recombinant retroviruses. J Virol 70:5701–5705

    CAS  PubMed  PubMed Central  Google Scholar 

  42. Christova Y, Adrain C, Bambrough P, Ibrahim A, Freeman M (2013) Mammalian iRhoms have distinct physiological functions including an essential role in TACE regulation. EMBO Rep 14:884–890

    Article  CAS  PubMed  PubMed Central  Google Scholar 

Download references

Acknowledgments

The authors record their thanks to previous members of the laboratory who contributed to the development of the study of early postinfection events in T. gondii immunity and pioneered the application of several of these techniques. The present work would also not have been possible without the contributions of the service facilities of the IGC, in particular the Animal Facility, supported by the research infrastructure Congento, project LISBOA-01-0145-FEDER-022170, the Transgenics Facility, and the Antibody facility, both supported by Fundação Calouste Gulenkian, the Advanced Imaging Unit, supported by the project PPBI-POCI-01-0145-FEDER-022122 and the Flow Cytometry Unit, supported by the project LISBOA-01-0145-FEDER-007654.

This work was supported by central funds of the Instituto Gulbenkian de Ciência, by the Sonderforschungsbereiche 670 and 680 and Schwerpunkt 1399 of the Deutsche Forschungsgemeinde. Joana Loureiro received funding from the European Union’s Horizon 2020 research and innovation program under the Marie Sklodowska-Curie Grant Agreement number 708694 entitled “Toxoplasma Sensing.”

Author contributions: Subheading 1, Jonathan Howard; Subheading 2, Catalina Alvarez and Ben Mueller; Subheadings 3 and 4, Ana Lina Rodrigues and Joana Loureiro; Subheading 5, Joana Loureiro; Subheading 6, Claudia Campos. All authors contributed to the preparation and editing of the entire manuscript.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Jonathan C. Howard .

Editor information

Editors and Affiliations

Appendices

Appendix 1: Generating Mammalian Cell Lines Expressing an HMGB1-GFP Necrosis Reporter (Joana Loureiro)

1.1 Generation of the Mouse HMGB1-GFP Construct

A murine HMGB1 cDNA (mHMGB1) with an in-frame C-terminal eGFP moiety (mHMGB1-GFP) was shown to be functionally similar to untagged, endogenous HMGB1 [24]. Murine messenger RNA (mRNA) obtained using the Qiagen RNeasy Kit (Cat No. 74104) from C57BL/6J MEFs induced for 24 h with murine IFNγ. RNA was DNase-treated (Thermo Fisher Scientific TURBO™ DNase, Cat No. AM2238) prior to cDNA synthesis (Invitrogen SuperScript III First-Strand Synthesis System, Cat No. 18080051). The 647 bp coding sequence of HMGB1 (Gene ID: 15289) was amplified from this cDNA (Herculase II Fusion DNA Polymerase, Agilent Technologies (Cat No. 600675) with primers JL1 and JL2 (see Table 3, below). After purification from an agarose gel (High Pure PCR Product Purification kit, Roche Cat. No. 11732668001), the amplicon was digested with EcoRI and SacII and cloned into EcoRI/SacII-digested Clontech’s pEGFP-N3 plasmid (now discontinued), ligated using Gibson Assembly® Master Mix (New England Biolabs, Cat. No. E2611S) and transformed into One Shot® Mach1™-T1R Chemically Competent E. coli. Positive colonies were identified by colony PCR and the sequence of the fusion gene in the pEGFP-N3 expression vector confirmed using Primer S1.

A functional HMGB1-GFP fusion protein should display a mostly nuclear localization in living mammalian cells, consistent with the nucleus-cytoplasm shuttling dynamics of endogenous untagged HMGB1 [22]. Indeed, cells with green-fluorescent nuclei were readily visualized and a protein of the expected size (~50–60 kDa)—GFP protein (238 aa) ~27 kDa; mHMGB1 (215 aa) ~20–30 kDa—was detected by western blot after transfection of pEGFP-N3 expressing the mHMGB1-GFP fusion construct into HEK293T cells (ATCC® CRL-3216™) (data not shown).

1.2 Generating the Retroviral mHGMB1-GFP Expression Construct

Because transfection efficiencies are often low in primary cells, we made a high-titer helper-free retroviral vector based on the pM6pBLAST system [25]: we used primers JL3 and JL4 (Table 3) to subclone the pEGFP-N3 mHMGB1-GFP cassette into pM6pBLAST. The ~1400 bp was digested with NcoI and NotI and cloned into Nco1/Not1 pM6pBLAST. A clone confirmed to contain the correct sequence (primers S2 and S3) showed the expected protein behavior both by microscopy and western blot following transfection of the viral plasmid DNA into HEK 293Tcells (data not shown).

Table 3 Primers used in generation of murine HMGB1-GFP constructs

The pM6pBLAST-based viral plasmid was used to produce high-titer retroviruses expressing the mHMGB1-GFP gene through cotransfection with a packaging plasmid into HEK 293Tcells, and the retroviral supernatants of HEK 293T cultures were used to transduce mouse or human fibroblasts. Both procedures are described in detail below.

1.3 Generation of Mammalian Cell Lines Expressing the HMGB1-GFP Reporter

In our experience, retrovirus-mediated transduction of mouse and human fibroblasts is desirable for several reasons. It is highly efficient: we typically obtain >70% of cells displaying green nuclei within 1 day of HMGB1-GFP-retrovirus-mediated transduction. Moreover, retroviral transduction results in stable integration of the HMGB1-GFP cassette into the genome, leading to long-term expression. Additionally, the Blasticidin-S resistance gene present in pM6pBLAST allows enrichment for HMGB1-GFP expression in cells in culture by means of antibiotic (blasticidin) selection.

HEK293T cells used for virus packaging must be of low passage and must be very carefully maintained such that they never reach confluence; the viruses must be packaged with a viral envelope appropriately pseudotyped to ensure mouse or human cell tropism (seeNote 1, below); the efficiency of virus transduction goes down exponentially with freeze-thawing of the virus-containing HEK293T cell culture supernatant, and therefore it is preferable to coordinate packaging and transduction so as to use freshly packaged virus.

1.3.1 Protocol for HMGB1-GFP Virus Production in HEK293T Cells

1.3.1.1 Material
  1. 1.

    DMEM supplemented with 10% heat-inactivated FCS, 1× NEAA, 1× penicillin–streptomycin, 1 mM sodium pyruvate, and 2 mM l-glutamine.

  2. 2.

    Cell culture incubator maintained at 37 °C with 10% CO2.

  3. 3.

    Sterile Flat bottom 6-well plates (Tissue culture-treated).

  4. 4.

    Pipette filter tips (20, 200, and 1000 μl).

  5. 5.

    Serological Pipettes (5 and 10 ml).

  6. 6.

    1.5 ml sterile microcentrifuge tubes.

  7. 7.

    ScreenFect®A plasmid transfection reagent and buffer (Incella Cat No. S-3001).

  8. 8.

    Retroviral DNA plasmid expressing murine HMGB1-GFP (fusion construct in pM6p.BLAST).

  9. 9.

    Packaging plasmid: pCL-Eco (mouse-pseudotyped envelope) plasmid or pCl-10A1 (human-pseudotyped envelope) [42].

  10. 10.

    15 ml Falcon tubes.

  11. 11.

    5 ml Plastic syringes.

  12. 12.

    Syringe Filters (Acrodisc® 25 mm Syringe filters with 0.45 μm Supor® Membrane).

1.3.1.2 Methods

Day −1: (Day before transfection) Plate cells at 750,000 cells/well in 6-well plates.

Day 0: (Day of transfection) Perform a visual assessment of confluence state of the cells. If cells are 65–75% confluent, you may proceed with the transfection procedure.

  1. 1.

    Replace medium on HEK293T cells—add 2 ml fresh FCS-containing DMEM (add medium VERY SLOWLY and CAREFULLY to the side walls of the wells, not directly to cells, or the cells will detach from the plate).

  2. 2.

    Make ScreenFect®A transfection mix and incubate for 20 min at RT as per the manufacturer’s instructions. Transfection mix: dilute SFA reagent in SFA dilution buffer in tube A and dilute DNA in SFA dilution buffer in tube B. DNA dilution in Tube B: For each well of a 6 well plate, dilute 2.5 μg of retroviral plasmid DNA + 1.25 μg of pCL-Eco (encoding mouse-pseudotyped viral envelope) or pCl-10A1 (encoding human-pseudotyped viral envelope) in 240 μl of ScreenFect®A dilution buffer.

  3. 3.

    After 20 min of RT incubation, add the SFA-DNA mix drop-wise and very carefully to HEK293T cells. Return cells to the incubator.

Day +1: Change medium—Replace medium (very carefully) on HEK293T cells.

Day +2: Collect 48 h virus supernatants.

  1. 1.

    Collect 48 h virus supernatants (sups) into 15 ml Falcon tubes and place sups at 4 °C Overnight.

  2. 2.

    Add 2 ml fresh serum-containing DMEM to each well and return HEK 293T cells to the incubator.

  3. 3.

    Optional: Plate target cells for infection after 18–24 h (Appendix 1.1.1.3).

Day +3: Collect 72 h virus supernatants.

  1. 1.

    Collect 72 h sups. Pool with 48 h virus supernatants.

  2. 2.

    Add polybrene to the (48 + 72) h supernatants to a final concentration of 8 μg/ml.

  3. 3.

    Filter through a 0.45 μm filter. Use supernatants immediately for infection of target cells (Appendix 1.1.1.3) or freeze aliquots at −20 °C. Note that the efficiency of virus transduction goes down exponentially with freeze-thawing.

1.3.2 Transduction of Target Cells with HMGB1-GFP-Expressing Retrovirus

Virus-mediated infection (transduction) by pCL-Eco or pCl-10A1-pseudotyped retrovirus has allowed us to generate HMGB1-GFP-expressing MEFs, DDCs, HFFs, and HeLa cells using the following protocol.

1.3.2.1 Material
  1. 1.

    DMEM supplemented with 10% heat-inactivated FCS, 1× NEAA, 1× penicillin–streptomycin, 1 mM sodium pyruvate, and 2 mM l-glutamine.

  2. 2.

    Cell culture incubator maintained at 37 °C with 10% CO2.

  3. 3.

    Sterile flat bottom 12-well plates (Tissue culture-treated).

  4. 4.

    Pipette filter tips (20, 200, and 1000 μl).

  5. 5.

    Serological Pipettes (5 and 10 ml).

  6. 6.

    Mouse- or human-pseudotyped virus supernatants (filtered and polybrene-supplemented).

  7. 7.

    Target cells plated the day prior to retroviral transduction on 12-well plates.

  8. 8.

    Tabletop centrifuge equipped with a high-capacity microplate rotor and plate holders (preferably one that is capable of maintaining cells at 37 °C during spin infection).

  9. 9.

    Blasticidin-S hydrochloride (CAS 3513-03-9).

1.3.2.2 Methods

Day −1: Plate murine/human target cells∗ such that target cells are no more than 50–60% confluent at the time of transduction (there must be room on the plate for cells to divide, since it is during cell division that the retrovirus can access the cell nucleus and integrate into the host DNA).

∗MEFs and DDCs are plated at 20000 or 15,000 cells/well, respectively, HFFs are plated at 32000 cells/well of a 12-well plate.

Day 0: Virus-mediated transduction of target cells.

  1. 1.

    Apply 37 °C-warm supernatants (filtered + supplemented with polybrene to 8 μg/ml) to target cells (typically 2 ml of virus supernatant/well of a 12-well plate).

  2. 2.

    Perform a spin-infection at 2200 rpm (~900 × g), at 37 °C, for 90 min. We use an Eppendorf centrifuge model 5810 R equipped with the A-4-81-MTP/Flex high-capacity microplate rotor (16.3 cm radius) with four plate holders.

  3. 3.

    After 90 min spin-infection, return cells to the 37 °C incubator.

Day +2 (48 h after target cell transduction): detach cells from plate and passage into 6-well plates and, if desired to increase the frequency of transduced cells, initiate selection with blasticidin-S HCl-containing DMEM. For antibiotic selection of pM6pBLAST-transduced MEFs we use 10 μg/ml of blasticidin; murine DDCs, human Hs27 fibroblasts and HeLa cells require higher blasticidin concentrations (15–20 μg/ml).

As early as 48 h after retrovirus transduction, expression of the mHGMB1-GFP fusion protein in mouse or human cells can be assessed. The predominantly nuclear localization and the electrophoretic mobility pattern of the mHGMB1-GFP reporter in C57BL/6J MEFs are shown in Fig. 9 (Table 3).

Fig. 9
figure 9

Expression of the mHMGB1-GFP retroviral reporter 48 h after transduction of C57BL/6J MEFs. (a) Merged phase contrast and green fluorescence images of transduced MEFs. (b) Western blot analysis of transduced (+) and mock-transduced (−) MEFs. Replica membranes were incubated with anti-human HMGB1 or anti-GFP antibodies. The higher molecular weight portions of the two membranes (100–150 kDa) were incubated with anti-human vinculin antibody as a loading control. The molecular mass of nuclear HMGB1 is ~25 kDa, whereas cytoplasmic and extracellular HMGB1 runs at ~30 kDa, largely due to acetylation. The multiple bands in the size range for the fusion protein likely reflect posttranslational modifications (acetylation, phosphorylation, and methylation) [25,26,27]

Appendix 2

See Fig. 10.

Fig. 10
figure 10

A fixation-permissive method for detecting necrosis by flow cytometry. (a) Schematic representation of fixable LIVE/DEAD® necrosis dyes. Live cells are dimly fluorescent, whereas dead cells, with a permeable membrane (indicated by the dashed line), display higher fluorescence. (b) Univariate histograms depicting differential staining by LIVE/DEAD® dye of C57BL/6J DDCs left untreated or exposed to a cell death inducer. In the absence of LIVE/DEAD® stain (“unstained” panel), untreated cells show no fluorescence. LIVE/DEAD® stained, untreated cells are mostly found in the LIVE cell gate (left peak) while cells exposed to high levels (100 mM) of hydrogen peroxide (H2O2) which damages the permeability barrier, are found in the DEAD cell gate (right peak) of the histogram. Cells treated with 10 μM staurosporine (Sta) to induce apoptosis remain mostly within the LIVE cell gate because the permeability barrier is not broken. About 20% of cells stain heavily with the dye, presumably due to secondary necrosis after Sta-induced apoptosis. Treatment with both agents was for 3 h before preparation for FACS. Samples were analyzed by flow cytometry using 633 nm excitation and ~661 nm emission (FACScalibur cytometer, FL4 channel). (c) Proportion of necrotic (LIVE/DEADhigh) cells in samples shown in b. (d) The morphology and PI staining profile of C57B/L6J DDCs shown in b are consistent with their LIVE/DEAD staining prolife. Shown are merged images obtained from the combination of the PI (red) channel and the phase contrast channel microscopy analysis of control, 100 mM H2O2- or 10 μM Sta-treated cells. Changes in cell morphology range from a flat shape (untreated controls) to a round shape with enlarged nuclei and expanded cytoplasm (necrotic cells, with PI+ nuclei) to a wrinkling, stellate shape (apoptotic cells, a few PI+ nuclei)

Appendix 3: Macro

/∗ IJ macro to create ROIs for parasite, PV etc... by GabyGMartins @Instituto Gulbenkian Ciencia, Advanced Imaging Facility - v0.2 2019-01-02 macro assumes images are loaded with bioformats in imageJ/FIJI as a composite hyperstack and reading of metadata expects pixel sizes to be properly scaled, as measurements are performed in microns and that chromatic shift has been corrected - if there is pixel-shift then measurements might not be accurate ∗/ roiManager("reset"); run("Select None"); run("Make Composite"); run("Channels Tool..."); //activates channel tool and gives user a chance to turn off unwanted channels (eg bright field) waitForUser("turn off unwanted channels then OK"); //asks user to switch to PV channel and store in variable PVchannel PVchannel = getNumber("Which channel do you use to detect the PV?", 3); Stack.setChannel(PVchannel); /∗ Proceed to identifying semi-automatically the PV: activates wand tool so user can click on parasite adjusting the "tolerance" of the wand the ROI matches the shape of the PV a default tolerance value of 4500 work with our images; a different default can be inserted below to accelerate to process if magic wand fails user can also change ROI selection tool manually and draw the parasite ∗/ setTool("zoom"); //activates zoom function so user can zoom in on parasite waitForUser("Click on parasite to zoom in then click OK"); setTool("wand"); run("Wand Tool...", "tolerance=4500 mode=4-connected"); // user can change the default tolerance here waitForUser("Click inside the parasite to select and double click on wand tool to adjust tolerance"); run("Enlarge...", "enlarge=0.1"); //user here can change the area of the ring around PV // stores the ROIcontaining the PV to ROImanager roiManager("Add"); roiManager("Select", 0); roiManager("rename", "PV") //create band of cytoplasm outside of parasite and stores it as cytosol in ROI manager run("Make Band...", "band=1"); //change thickness of band here roiManager("Add"); roiManager("Select", 1); roiManager("rename", "cytosol") // create band of 0.8 μm representing the parasite's cortex and call it "PVM" roiManager("Select", 0); run("Enlarge...", "enlarge=-0.8"); run("Make Band...", "band=0.8"); roiManager("Add"); roiManager("Select", 2); roiManager("rename", "PVM") // prepare ROImanager for measurements roiManager("Select", newArray(0,1,2)); roiManager("Show All"); //activates several measurements and then stores measurements into a results table run("Set Measurements...", "area mean standard modal min shape feret's integrated stack display redirect=None decimal=3");roiManager("multi-measure measure_all");

Rights and permissions

Reprints and permissions

Copyright information

© 2020 Springer Science+Business Media, LLC, part of Springer Nature

About this protocol

Check for updates. Verify currency and authenticity via CrossMark

Cite this protocol

Alvarez, C., Campos, A.C., Howard, J.C., Loureiro, J., Müller, U.B., Rodrigues, A.L. (2020). Methods for the Measurement of Early Events in Toxoplasma gondii Immunity in Mouse Cells. In: Tonkin, C. (eds) Toxoplasma gondii. Methods in Molecular Biology, vol 2071. Humana, New York, NY. https://doi.org/10.1007/978-1-4939-9857-9_20

Download citation

  • DOI: https://doi.org/10.1007/978-1-4939-9857-9_20

  • Published:

  • Publisher Name: Humana, New York, NY

  • Print ISBN: 978-1-4939-9856-2

  • Online ISBN: 978-1-4939-9857-9

  • eBook Packages: Springer Protocols

Publish with us

Policies and ethics