This paper is available on line at ...cabm-ms.cabm.rutgers.edu/Classification of subcellular...

14
Classification of Subcellular Location by Comparative Proteomic Analysis of Native and Density-shifted Lysosomes* S Maria Cecilia Della Valle‡§, David E. Sleat‡§, Haiyan Zheng‡, Dirk F. Moore¶, Michel Jadot, and Peter Lobel‡§** One approach to the functional characterization of the lysosome lies in the use of proteomic methods to identify proteins in subcellular fractions enriched for this organ- elle. However, distinguishing between true lysosomal res- idents and proteins from other cofractionating organelles is challenging. To this end, we implemented a quantitative mass spectrometry approach based on the selective de- crease in the buoyant density of liver lysosomes that oc- curs when animals are treated with Triton-WR1339. Liver lysosome-enriched preparations from control and treated rats were fractionated by isopycnic sucrose density gra- dient centrifugation. Tryptic peptides derived from gradi- ent fractions were reacted with isobaric tag for relative and absolute quantitation eight-plex labeling reagents and analyzed by two-dimensional liquid chromatography matrix-assisted laser desorption ionization time-of-flight MS. Reporter ion intensities were used to generate rela- tive protein distribution profiles across both types of gra- dients. A distribution index was calculated for each iden- tified protein and used to determine a probability of lysosomal residence by quadratic discriminant analysis. This analysis suggests that several proteins assigned to the lysosome in other proteomics studies are not true lysosomal residents. Conversely, results support lyso- somal residency for other proteins that are either not or only tentatively assigned to this location. The density shift for two proteins, Cu/Zn superoxide dismutase and ATP-bind- ing cassette subfamily B (MDR/TAP) member 6, was cor- roborated by quantitative Western blotting. Additional bal- ance sheet analyses on differential centrifugation fractions revealed that Cu/Zn superoxide dismutase is predominantly cytosolic with a secondary lysosomal localization whereas ATP-binding cassette subfamily B (MDR/TAP) member 6 is predominantly lysosomal. These results establish a quanti- tative mass spectrometric/subcellular fractionation ap- proach for identification of lysosomal proteins and under- score the necessity of balance sheet analysis for localization studies. Molecular & Cellular Proteomics 10: 10.1074/mcp.M110.006403, 1–14, 2011. Lysosomes are membrane-delimited organelles that are re- sponsible for the degradation of macromolecules delivered via various cellular pathways including endocytosis, phagocytosis, and autophagy. The catabolic function of the lysosome reflects the concerted action of numerous hydrolases that break down large molecules into simpler components that can be reutilized by the cell. The biomedical importance of this organelle is un- derscored by the devastating effects of deficiencies in its com- ponents, with more than 40 different lysosomal storage dis- eases that result from mutations in genes encoding lysosomal proteins (1). Alterations in lysosomal function have also been linked with more widespread human disorders, including cancer (2) and neurodegenerative disorders such as Alzheimer and Parkinson diseases (3–5), thus there is a growing interest in understanding the function of this organelle. One approach to this lies in characterizing its proteome. To date, more than 60 different soluble lumenal proteins and numerous membrane proteins have been associated with the lysosomal compart- ment. However, new lysosomal proteins continue to be discov- ered and it is likely that more remain to be classified as such. In recent years, two general protein identification ap- proaches have been used to investigate the lysosomal pro- teome (reviewed in (1, 6 – 8)). One approach focuses on solu- ble proteins, using affinity purification to isolate proteins targeted to the lysosome via the mannose 6-phosphate (Man6-P) 1 pathway. The other approach uses subcellular From the ‡Center for Advanced Biotechnology and Medicine and §Department of Pharmacology, University of Medicine and Dentistry of New Jersey, Robert Wood Johnson Medical School, Piscataway, NJ 08854, ¶Department of Biostatistics, University of Medicine and Dentistry of New Jersey, School of Public Health, Piscataway, NJ 08854, and Laboratoire de Chimie Physiologique, Unite ´ de Recher- che en Physiologie Mole ´ culaire, Faculte ´ s Universitaires Notre-Dame de la Paix, 61 rue de Bruxelles, 5000-Namur, Belgium Received November 10, 2010, and in revised form, January 19, 2011 Published, MCP Papers in Press, January 20, 2011, DOI 10.1074/ mcp.M110.006403 1 The abbreviations used are: ACN, acetonitrile; ABCB6, ATP-bind- ing cassette subfamily B (MDR/TAP) member 6; ACP2, classical lysosomal acid phosphatase; ACP5, tartrate-resistant acid phospha- tase; APCS, amyloid P-component, serum; -gal, -galactosidase; CAT, catalase; GPM, Global Proteome Machine; iTRAQ, isobaric tag for relative and absolute quantitation; L, light mitochondrial; M, heavy mitochondrial; Man 6-P, mannose 6-phosphate; N, nuclear; P, micro- somal; PLBD1, phospholipase B domain containing 1; QDA, quad- ratic discriminant analysis; S, cytosolic; SCX, strong cation exchange; SOD1, Cu/Zn superoxide dismutase 1; UOX, urate oxidase. Technological Innovation and Resources © 2011 by The American Society for Biochemistry and Molecular Biology, Inc. This paper is available on line at http://www.mcponline.org Molecular & Cellular Proteomics 10.4 10.1074/mcp.M110.006403–1

Transcript of This paper is available on line at ...cabm-ms.cabm.rutgers.edu/Classification of subcellular...

Page 1: This paper is available on line at ...cabm-ms.cabm.rutgers.edu/Classification of subcellular location by... · Classification of Subcellular Location by Comparative Proteomic Analysis

Classification of Subcellular Location byComparative Proteomic Analysis of Native andDensity-shifted Lysosomes*□S

Maria Cecilia Della Valle‡§, David E. Sleat‡§, Haiyan Zheng‡, Dirk F. Moore¶,Michel Jadot�, and Peter Lobel‡§**

One approach to the functional characterization of thelysosome lies in the use of proteomic methods to identifyproteins in subcellular fractions enriched for this organ-elle. However, distinguishing between true lysosomal res-idents and proteins from other cofractionating organellesis challenging. To this end, we implemented a quantitativemass spectrometry approach based on the selective de-crease in the buoyant density of liver lysosomes that oc-curs when animals are treated with Triton-WR1339. Liverlysosome-enriched preparations from control and treatedrats were fractionated by isopycnic sucrose density gra-dient centrifugation. Tryptic peptides derived from gradi-ent fractions were reacted with isobaric tag for relativeand absolute quantitation eight-plex labeling reagentsand analyzed by two-dimensional liquid chromatographymatrix-assisted laser desorption ionization time-of-flightMS. Reporter ion intensities were used to generate rela-tive protein distribution profiles across both types of gra-dients. A distribution index was calculated for each iden-tified protein and used to determine a probability oflysosomal residence by quadratic discriminant analysis.This analysis suggests that several proteins assigned tothe lysosome in other proteomics studies are not truelysosomal residents. Conversely, results support lyso-somal residency for other proteins that are either not or onlytentatively assigned to this location. The density shift fortwo proteins, Cu/Zn superoxide dismutase and ATP-bind-ing cassette subfamily B (MDR/TAP) member 6, was cor-roborated by quantitative Western blotting. Additional bal-ance sheet analyses on differential centrifugation fractionsrevealed that Cu/Zn superoxide dismutase is predominantlycytosolic with a secondary lysosomal localization whereasATP-binding cassette subfamily B (MDR/TAP) member 6 ispredominantly lysosomal. These results establish a quanti-

tative mass spectrometric/subcellular fractionation ap-proach for identification of lysosomal proteins and under-score the necessity of balance sheet analysis forlocalization studies. Molecular & Cellular Proteomics 10:10.1074/mcp.M110.006403, 1–14, 2011.

Lysosomes are membrane-delimited organelles that are re-sponsible for the degradation of macromolecules delivered viavarious cellular pathways including endocytosis, phagocytosis,and autophagy. The catabolic function of the lysosome reflectsthe concerted action of numerous hydrolases that break downlarge molecules into simpler components that can be reutilizedby the cell. The biomedical importance of this organelle is un-derscored by the devastating effects of deficiencies in its com-ponents, with more than 40 different lysosomal storage dis-eases that result from mutations in genes encoding lysosomalproteins (1). Alterations in lysosomal function have also beenlinked with more widespread human disorders, including cancer(2) and neurodegenerative disorders such as Alzheimer andParkinson diseases (3–5), thus there is a growing interest inunderstanding the function of this organelle. One approach tothis lies in characterizing its proteome. To date, more than 60different soluble lumenal proteins and numerous membraneproteins have been associated with the lysosomal compart-ment. However, new lysosomal proteins continue to be discov-ered and it is likely that more remain to be classified as such.

In recent years, two general protein identification ap-proaches have been used to investigate the lysosomal pro-teome (reviewed in (1, 6–8)). One approach focuses on solu-ble proteins, using affinity purification to isolate proteinstargeted to the lysosome via the mannose 6-phosphate(Man6-P)1 pathway. The other approach uses subcellular

From the ‡Center for Advanced Biotechnology and Medicine and§Department of Pharmacology, University of Medicine and Dentistryof New Jersey, Robert Wood Johnson Medical School, Piscataway,NJ 08854, ¶Department of Biostatistics, University of Medicine andDentistry of New Jersey, School of Public Health, Piscataway, NJ08854, and �Laboratoire de Chimie Physiologique, Unite de Recher-che en Physiologie Moleculaire, Facultes Universitaires Notre-Damede la Paix, 61 rue de Bruxelles, 5000-Namur, Belgium

Received November 10, 2010, and in revised form, January 19, 2011Published, MCP Papers in Press, January 20, 2011, DOI 10.1074/

mcp.M110.006403

1 The abbreviations used are: ACN, acetonitrile; ABCB6, ATP-bind-ing cassette subfamily B (MDR/TAP) member 6; ACP2, classicallysosomal acid phosphatase; ACP5, tartrate-resistant acid phospha-tase; APCS, amyloid P-component, serum; �-gal, �-galactosidase;CAT, catalase; GPM, Global Proteome Machine; iTRAQ, isobaric tagfor relative and absolute quantitation; L, light mitochondrial; M, heavymitochondrial; Man 6-P, mannose 6-phosphate; N, nuclear; P, micro-somal; PLBD1, phospholipase B domain containing 1; QDA, quad-ratic discriminant analysis; S, cytosolic; SCX, strong cation exchange;SOD1, Cu/Zn superoxide dismutase 1; UOX, urate oxidase.

Technological Innovation and Resources© 2011 by The American Society for Biochemistry and Molecular Biology, Inc.This paper is available on line at http://www.mcponline.org

Molecular & Cellular Proteomics 10.4 10.1074/mcp.M110.006403–1

Page 2: This paper is available on line at ...cabm-ms.cabm.rutgers.edu/Classification of subcellular location by... · Classification of Subcellular Location by Comparative Proteomic Analysis

fractionation to isolate membranes enriched in lysosomal pro-teins. Coupled with MS-based methods for protein identifica-tion, both approaches have been extremely useful in identify-ing many candidate lysosomal proteins. However, validationof lysosomal localization remains a critical step and is con-ducted on a protein-by-protein basis (reviewed in (1, 7, 8)).

A general framework for assignment of proteins to thelysosome using analytical subcellular fractionation was devel-oped over 50 years ago (9) and these principles remain highlyapplicable today. In the first step, a tissue homogenate isfractionated using differential centrifugation and balancesheets are used to monitor the distribution of a protein ofinterest among all fractions and to compare this with thedistribution of known organelle marker proteins. In subse-quent steps, a differential centrifugation fraction enriched inlysosomes is subjected to further fractionation methods, typ-ically isopycnic density gradient centrifugation, and the dis-tribution of the protein of interest again is compared with thatof known markers. If the protein cofractionates with bona fidelysosomal markers, this is consistent with a lysosomal resi-dence, and the more independent fractionation methods usedthat demonstrate this behavior, the greater the confidence ofthis assignment. Conversely, if a protein does not cofraction-ate with lysosomal markers, this demonstrates that it is not alysosomal resident. In some cases, proteins or activities havecomposite distributions with reference to known markers andthese can be quantitatively assigned to multiple locations (10).

In many fractionation schemes, the sedimentation coeffi-cients and/or buoyant densities of lysosomes overlap or areclose to those of other organelles, in particular mitochondriaand peroxisomes. Fortunately, there are methods to specifi-cally change the biophysical properties of the lysosome toalter its migration in a density gradient, thus providing anadditional level of resolution. In particular, treatment of ro-dents with Triton WR-1339 results in a marked decrease in thebuoyant density of liver lysosomes (converting them to “trito-somes”) but not of other organelles (11). This density shift is aspecific hallmark of lysosomal proteins that has been used inthe case-by-case verification of lysosomal candidates identi-fied in proteomics experiments (12–17).

In this study, we have combined the selectivity of the lyso-somal density shift with the protein identification and quantifi-cation capabilities of tandem mass-spectrometry on isobariclabeled samples as a step toward concomitant discovery andvalidation of lysosomal candidates. We assign a lysosomal lo-calization to several proteins that have not previously beenlocalized or whose location is controversial. Extensions of thisapproach have significant potential toward expanding the un-derstanding of lysosome complexity and function in biology andmedicine.

EXPERIMENTAL PROCEDURES

Subcellular Fractionation—All experiments and procedures involvinglive animals were conducted in compliance with approved Institutional

Animal Care and Use Committee protocols. Liver homogenates fromcontrol or Triton WR-1339 injected adult male Wistar rats (250–300 g)were prepared as described (18). All subsequent operations were con-ducted at 0 to 4 °C. Fractionation of subcellular organelles by differentialcentrifugation to obtain nuclear (N), heavy mitochondrial (M), light mi-tochondrial (L), microsomal (P) and cytosolic (S) fractions was as de-scribed (10). Isopycnic centrifugation (19) was performed with the fol-lowing modifications. For density gradients of control and triton-treatedrat liver, an appropriate amount of an ML fraction (combined heavy andlight mitochondrial) or an L fraction, was layered onto the top of linearsucrose gradients (density limits � 1.06–1.26 g/cm3) and centrifugedfor 150 min at 39,000 rpm in either a SW55 Ti or SW50 rotor. Twelve or13 fractions were collected following centrifugation using a tube slicerand densities were measured by refractometry. Samples were stored at�80 °C prior to further analysis.

Enzyme and Protein Assays—�-galactosidase and �-glucuroni-dase were measured using 4-methylumbelliferyl substrates (20). Cyto-chrome C oxidase was measured with a kinetic method (18, 21) at roomtemperature using 22 �M cytochrome C reduced with sodium dithioniteas a substrate in 30 mM sodium phosphate pH 7.4 containing 1 mM

EDTA. The decrease in absorbance at 550 nm was monitored during thefirst 2 min and is proportional to the activity of cytochrome C oxidase.Catalase was measured as described (22). Protein concentrations weremeasured using Advanced Protein Assay reagent (Cytoskeleton Inc.,Denver, CO) using bovine serum albumin standards.

Protein Microchemistry for Mass Spectrometry—Equivalent pro-portions of individual fractions were pooled as shown in Fig. 1 to yielda total of 45 �g protein per pool and adjusted to the same sucroseconcentration and final volume (typically 400 �l). The exception wasthat, because of sample limitations, pool C1 in Experiment II con-tained 27 �g protein. Two purified histidine-tagged bacterial proteins(Northeast Structural Genomics Consortium target identification num-bers DrR57 and GmR40, kindly provided by Dr. Guy Montelione) wereadded as internal standards (0.3 �g each protein). For solution di-gests, samples were then adjusted to contain 0.1% octyl-�-gluco-side, 6 M guanidine hydrochloride, and 10 mM dithiotreitol in a finalvolume of 800 �l, incubated for 30 min at 60 °C, cooled to 20 °C thencarbamidomethylated with 20 mM iodoacetamide for 1 h. Sampleswere buffer exchanged to 50 mM ammonium bicarbonate, pH 7.8 byultrafiltration using Vivaspin 500 10,000 MWCO concentrators(Vivaproducts, Littleton, MA). Modified porcine trypsin (Promega,Madison, WI) was added to retentates at a 1:50 w/w ratio trypsin:substrate and samples were incubated at 37 °C for 16 h. Peptideswere then recovered in the filtrate by centrifugal ultrafiltration at14,000 � g followed by two cycles of addition of 150 �l 50 mM

ammonium bicarbonate and ultrafiltration. These filtrates werepooled, dried using a vacuum centrifuge, and residual ammoniumbicarbonate was removed by three cycles of resuspension in 200 �l50% methanol and drying. Dried samples were stored at �80 °C. Forin-gel tryptic digestion, samples were dissolved with 1� lithium do-decyl sulfate gel loading buffer (Invitrogen) containing 10 mM dithio-treitol, heated for 10 min at 70 °C, and loaded into a 2 mm thickpolyacrylamide gel with large (14 mm wide � 20 mm deep) wells. Thegel consisted of 1 cm of 10% acrylamide segment cast on top of a20% acrylamide spacer that served to retain proteins in the 10%acrylamide segment. Electrophoresis was conducted until prestainedstandards in a separate well completely entered the gel. Gels werefixed, stained with colloidal Coomassie blue and the 10% acrylamidesegment containing each sample excised for subsequent reduction,carbamidomethylation and in-gel tryptic digestion (23). Peptides wereextracted from the gel pieces using 60% acetonitrile (ACN)/5% formicacid, dried using a vacuum concentrator, and ammonium bicarbonateremoved as described above.

Subcellular Fractionation and Lysosomal Proteomics

10.1074/mcp.M110.006403–2 Molecular & Cellular Proteomics 10.4

Page 3: This paper is available on line at ...cabm-ms.cabm.rutgers.edu/Classification of subcellular location by... · Classification of Subcellular Location by Comparative Proteomic Analysis

Isobaric Tag for Relative and Absolute Quantitation (iTRAQ) Label-ing—iTRAQ 8-plex labeling (24) was performed according to manu-facturer’s protocols (Applied Biosystems, Carlsbad, CA). Briefly, driedtryptic digests were dissolved in 20 to 25 �l of 0.5 M triethylammo-nium and iTRAQ 8-plex reagents dissolved in 50 �l of isopropanolwere added to the samples that were incubated for 2 h at 20 °C.Labels used for different samples are given in supplemental Fig. S1.Labeling was verified prior to pooling either by LC-MS/MS analysisusing an LTQ (Thermo Fisher) or by matrix-assisted laser desorptionionization tandem MS (MALDI-MS/MS) using an ABI 4800 (AppliedBiosystems). Most (�96%) of the peptides contained the iTRAQmodification, indicating that the labeling reactions were essentiallycomplete. Samples labeled with different iTRAQ reagents were driedusing a vacuum centrifuge, dissolved in 50% methanol, combined,vacuum-dried and stored at �80 °C until use.

Two-Dimensional Peptide Separation—Tryptic digests were dis-solved in 100 �l 0.1% trifluoroacetic acid (TFA) and loaded onto apipette tip column containing C18 resin (SPEC, Varian, Lake Forest,CA) equilibrated with 0.1% TFA. The column was washed with 600 �l0.1% TFA and peptides were sequentially eluted with 100 �l of 30%ACN/0.1% TFA, 100 �l of 50% ACN/0.1% TFA, and 100 �l of 80%ACN/0.1% TFA. The procedure was repeated with the column flowthrough as above. The eluates were pooled, vacuum-dried, and usedfor strong cation exchange (SCX) chromatography. Briefly, peptideswere dissolved in 180 �l of buffer A (5 mM KH2PO4 and 25% ACN, pH3.0) and applied to a 1 mm � 150 mm, 5 �m, 300Å polysulfoethyl Acolumn (PolyLC Inc, Columbia, MD). Chromatography was con-ducted at a flow rate of 40 �l/min using an Ultimate LC system(Dionex, Sunnyvale, CA) equipped with an UV absorbance monitor.The column was washed with Buffer A for 5 min and developed witha gradient of 0–100% Buffer B (5 mM KH2PO4, 25% ACN, and 400 mM

KCl, pH 3.0) in 30 min followed by 100% Buffer B for 5 min. Fractionswere collected at 2 min intervals and vacuum-dried. SCX fractionswere dissolved in 0.1% TFA, combined based on absorbance mea-surements to yield seven SCX pooled fractions that contained roughlyequivalent amounts of peptides. Each of these was further fraction-ated by reverse phase high performance liquid chromatography andthe eluate deposited onto a MALDI plate. Briefly, samples wereloaded onto a 300 �m � 5 mm C18 trap column (Dionex, Sunnyvale,CA) and washed with 0.1% TFA for 5 min at a flow rate of 25 �l/min.The flow was reversed and the trap column brought in line with a 75�m � 12 cm column packed in-house with Magic C18AQ (3 �m, 200Å; Michrom BioResources Inc., Auburn, CA). The columns were de-veloped using a linear gradient of 2 to 50% ACN in 0.1% TFA at a flowrate of 250 nl/min for 120 min. The outflow was mixed with a 2 �l/minstream of 50% ACN/0.1% TFA solution using a low dead volume teeand fractions were deposited onto a 384-well MALDI plate using aProbot plate spotter (Dionex) (15 s/fraction). Matrix (2.5 mg/ml recrys-tallized �-cyano-4-hydroxy cinnamic acid in 50% ACN/0.1% TFA)was manually added to each spot before MALDI-MS/MS analysis.

Tandem Mass Spectrometry—Tandem MALDI mass spectrometryof iTRAQ labeled samples was conducted using an Applied Biosys-tems 4800 MALDI-TOF/TOF. MS spectra were acquired in a windowof m/z 800–4000 in positive ion reflectron mode. The 10 most abun-dant precursor ions with a signal to noise ratio greater than 50 wereselected for top-down MS/MS scans, excluding identical precursorions present in adjacent spots from a given LC-MALDI run. Precursorions were selected at a relative resolution of 200 full width at halfmaximum and MS/MS was conducted using a collision energy of 1 kV(Experiments Ia and Ib) or 2 kV (Experiment II) in positive ion mode,accumulating 2000 laser shots per spectra. Peaks with a minimumsignal to noise ratio of 10 and a mass range of 60 Da to 20 Da belowthe precursor ion mass were exported from the Applied Biosystems4000 Series database into mgf files using the TS2 Mascot utility.

Database Searching—Searches were conducted using a local im-plementation of the Global Proteome Machine (GPM) XE Managerversion 2.2.1 (Beavis Informatics Ltd., Winnipeg, Canada), which usesX!Tandem Tornado 2010.01.01 to assign spectral data (25). All mgffiles from the MALDI-TOF/TOF experiments were merged and usedto search a combined database consisting of the ENSEMBL ratand mouse proteomes (Rattus_norvegicus.RGSC3.4.58.pep.all andMus_musculus.NCBIM37.58.pep.all), the trypsin and dust/contami-nant proteins abstracted from the GPM cRAP database and the tworecombinant bacterial fusion proteins used as internal standards (seebelow). Search parameters specified a precursor ion mass error of100 ppm and a fragment mass error of 0.4 Da with a minimum numberof 5 MS/MS fragments. Cysteine carbamidomethylation and iTRAQlabeling were constant modifications, oxidation of methionine was avariable modification, and one missed cleavage was allowed duringthe preliminary model development. The threshold used for modelrefinement was a peptide expectation score of 0.01. During refine-ment, deamidation at asparagine and glutamine, and oxidation atmethionine residues were allowed. Proteins were mapped back tocorresponding genes and only those with at least two independentpeptides and GPM expectation scores of 10�5 or better were ac-cepted as valid identifications.

Data Analysis—Intensities associated with reporter ions (monoiso-topic mass � 0.06 Da for each reporter) were extracted from mgf filesusing a custom PERL script to generate a list of spectra and associ-ated information. Matrix functions in Microsoft Excel were used toadjust iTRAQ reporter ion peak areas for crossover using correctionfactors supplied by the vendor. This data set was merged with anExcel spreadsheet containing a list of all spectra identified in data-base searching (see above). Spectra were used for quantification onlyif they met all of the following criteria: (1) the mean iTRAQ reporter ionintensity for both the control and Triton WR-1339 treated gradientfractions were �300 (see “Results”); (2) the assigned peptide con-tained a maximum of one missed tryptic cleavage sites; and (3) theassigned peptide was either fully tryptic or semitryptic. Gene prod-ucts required at least three quantifiable spectra for inclusion in theanalysis. Internal standards from each experiment were used to gen-erate correction factors that were applied to the ion intensity data toadjust for differential losses and labeling efficiencies (see “Results”).

Calculation of Distribution Profiles and Statistical Analysis—EachMS/MS spectrum included in the analysis contains eight reporter ionintensities that correspond to the four different samples from each ofthe two different sucrose density gradients analyzed in a given ex-periment. Each spectrum was used to create two sets of normalizedvalues, where the reporter ion associated with a given gradient frac-tion was divided by the sum of the four reporter ions for that gradient.Normalized ion intensities for all spectra assigned to a given proteinwere averaged and used for plotting distribution profiles and forquadratic discriminant analysis (QDA) (26). To account for underlyingdata quality, we applied a parametric bootstrap analysis to refine theQDA (see “Results”) using R version 2.11.0, with contour lines thatseparate the different classes generated using the “klaR” library. A setof rodent genes encoding established lysosomal matrix and mem-brane proteins or strong candidates compiled from Tables I–III ofLubke et al. (1), Table II of Sleat et al. (23), and Tables II and III ofSchroeder et al. (8) (“curated lysosomal”) (supplemental Table S1),was used as a starting point for predicting new lysosomal candidates.

Western Blotting—Samples were fractionated by SDS-PAGE onprecast NuPAGE Novex 10% bis-Tris or 3–8% Tris-acetate midigels(Invitrogen Corporation, Carlsbad, CA) under reducing conditions andtransferred to nitrocellulose. For quantitative analysis, different loadvolumes were used to ensure that the signal was in the linear range.Membranes were probed with either a sheep antisera against super-oxide dismutase, Cu/Zn enzyme (# 574597, lot D00047669, Calbi-

Subcellular Fractionation and Lysosomal Proteomics

Molecular & Cellular Proteomics 10.4 10.1074/mcp.M110.006403–3

Page 4: This paper is available on line at ...cabm-ms.cabm.rutgers.edu/Classification of subcellular location by... · Classification of Subcellular Location by Comparative Proteomic Analysis

ochem, San Diego, CA) or a rabbit antisera against ABCB6 (# LS-C19004, lot 8007, LifeSpan BioSciences, Inc., Seattle, WA) andappropriate iodinated secondary antibodies. Signal was visualizedand quantified by phosphorimager analysis using a Typhoon 9400scanner and ImageQuant 5.2 software (GE Healthcare Bio-Sciences).

RESULTS

Experimental Workflow—The aim of this study was to com-bine classical approaches to subcellular fractionation withquantitative mass-spectrometry to identify lysosomal proteinswith high confidence. The first step was to prepare lightmitochondrial (L) differential centrifugation fractions (whichare �4–10-fold enriched in lysosomes compared with thehomogenate as determined by �-galactosidase assay) fromindividual control and Triton WR-1339 treated rats and furtherfractionate these using sucrose density gradient centrifuga-tion. Two independent pairs of control and treated animalswere analyzed. As expected, gradients from the treated rats(“Triton gradients”) contained a peak of the lysosomal markeractivity �-galactosidase in the lower density fractions (Fig. 1,Top Panels), whereas in gradients from control animals (“con-trol gradients”), this marker was in the higher density fractions(Fig. 1, Bottom Panels). The distribution of the mitochondrialmarker cytochrome oxidase, the peroxisomal marker cata-lase, and total protein were essentially unchanged by thetreatment (Fig. 1).

Fig. 2 shows the overall workflow for the mass spectrom-etry experiments that were used to analyze each pair of con-trol and treated samples and additional details are provided insupplemental Fig. S1. Although each gradient typically is di-vided into 12–13 fractions for marker enzyme analysis, thisnumber of samples is impractical for iTRAQ analysis. Thus,

each gradient was divided into four fractions (C1–4 and T1–4

for the control and Triton gradients, respectively: see Fig. 1),resulting in a set of eight fractions. To each of these fractions,recombinant bacterial fusion proteins were added as internalstandards, then they were digested with trypsin and eachlabeled with a different iTRAQ 8-plex reagent. The eightiTRAQ-labeled samples were pooled and the peptides furtherfractionated by two-dimensional liquid chromatography andanalyzed using a MALDI-TOF/TOF mass spectrometer. Thefirst set of samples was used for two experiments. In Exper-iment Ia, samples were reduced, alkylated, and digested withtrypsin in solution, whereas in Experiment Ib, these stepswere conducted on proteins run into a polyacrylamide gel.Based on the results from this analysis (see below), the sec-ond set of samples were processed in-gel (Experiment II). Inaddition to representing samples derived from different ani-mals, Experiment I and II differed in terms of which iTRAQ8-plex reagent was used to label fractions from the differentgradient fractions and on the collision energies used forMS/MS (supplemental Fig. S1).

Evaluation of Sample Workflows and Data Quality UsingInternal Standards—The internal standards allowed evalua-tion of potential unequal losses and labeling efficienciesamong the eight individual fractions before the iTRAQ labeledsamples were combined. To determine whether either sourceof error was significant, for each spectrum, iTRAQ reporter ionintensity was normalized to the sum of all eight reporter ionintensities. If all eight samples had exactly the same labeling

FIG. 1. Distribution of organelle markers and protein in sucrosedensity gradient fractions. Differential centrifugation L fractions pre-pared from livers of control or Triton WR-1339 treated rats werefractionated by isopycnic sucrose density gradient centrifugation.Each gradient represents an individual animal. Indicated fractionswere combined following marker analysis as shown to create fourpools per gradient.

FIG. 2. Workflow for sample processing and iTRAQ labeling.Liver homogenates were processed and used for quantitative massspectrometry as described in text.

Subcellular Fractionation and Lysosomal Proteomics

10.1074/mcp.M110.006403–4 Molecular & Cellular Proteomics 10.4

Page 5: This paper is available on line at ...cabm-ms.cabm.rutgers.edu/Classification of subcellular location by... · Classification of Subcellular Location by Comparative Proteomic Analysis

efficiency and losses, then the normalized intensity for alleight reporters would be exactly 0.125. Conversely, differencesin normalized intensities for the internal standards would indi-cate differential losses or unequal labeling efficiencies. Compar-ison of Experiments Ia and Ib indicated slightly less variability forthe in-gel digested samples compared with the solution digests(supplemental Fig. S2). Digestion format per se was not neces-sarily related to variability in the data but given that both wereacceptable, we proceeded with in-gel digests for Experiment II.This analysis also allowed for the generation of correction fac-tors that we applied to each reporter ion in each experiment toaccount for variability associated with sample processing. Fol-lowing correction, data from Experiments Ia and Ib were pooledand used for subsequent analysis.

We also examined the effect of reporter ion intensity on signalvariance ((supplemental Fig. S3). As expected, the coefficient ofvariation for each spectrum associated with the internal stand-ards tends to be inversely proportional to the average iTRAQintensity. Based upon these results, we only used spectra wherethe average reporter ion intensity was �300.

Protein Identification and Quantitation—The first pass da-tabase search using the GPM resulted in assignment of 8913spectra to 1273 rodent gene products (excluding internalstandards and contaminants) with a peptide false discoveryrate of 0.77% (supplemental Tables S2 and S3). These as-signments included 56 curated lysosomal proteins (see Meth-ods). Filtering data to include only proteins with a �10�5

chance of a stochastic match that also had at least twounique peptides reduced this to 657 gene products, 41 beingcurated lysosomal. Further filtering to include gene productswith at least three quantifiable spectra (see Methods) reducedthis to 7590 quantifiable spectra (5588 in Experiment I, 2002 inExperiment II) assigned to 545 different gene products, 38being curated lysosomal. For the latter, 292 spectra in Exper-iment I were assigned to 37 proteins whereas 184 spectra inExperiment II were assigned to 38 proteins.

For each experiment, reporter ion intensities from eachspectrum were first corrected for internal standard recovery,then normalized and used to create distribution profiles for thecontrol and triton gradients. Profiles from all quantifiablespectra assigned to a given gene product were averaged tocreate mean distribution profiles. Fig. 3 shows profiles ofrelatively abundant proteins that have been assigned to lyso-somes, mitochondria, peroxisomes, and endoplasmic reticu-lum as well as other miscellaneous proteins.

The mean distribution profiles for the lysosomal proteasecathepsin D (CSTD), classical lysosomal acid phosphatase(ACP2), and the lysosomal membrane protein LAMP2 aresimilar to each other and to that of the relative specific enzy-matic activity of lysosomal �-galactosidase (�-gal). The con-trol gradient fractions have similar relative intensities or spe-cific activities (compare fractions C1 to C4). In contrast, thereis evident enrichment in the lower density fractions of thetriton gradients (compare fractions T1 and T2 with T3 and T4).

Thus, the iTRAQ labeling strategy can clearly detect the den-sity shift characteristic of lysosomal proteins.

Analysis of other types of proteins indicated that their dis-tributions in the density gradients were largely unaffected by

FIG. 3. Distribution of different classes of proteins in sucrosedensity gradient pooled fractions. Relative intensities were cal-culated as described in Methods. Error bars show the 95% confi-dence intervals calculated using Prism5.03 (GraphPad Software,Inc). The profile for �-galactosidase is based on enzyme activityand protein measurements conducted following fractions werepooled. Note that the specific reporter ion intensities or activity arenormalized to protein levels. Thus, even though most of the lyso-somal proteins sediment in the denser fractions in the controlgradients as shown in Fig. 1, these fractions also contain the bulkof the protein, resulting in similar relative specific activities amongthe pooled fractions.

Subcellular Fractionation and Lysosomal Proteomics

Molecular & Cellular Proteomics 10.4 10.1074/mcp.M110.006403–5

Page 6: This paper is available on line at ...cabm-ms.cabm.rutgers.edu/Classification of subcellular location by... · Classification of Subcellular Location by Comparative Proteomic Analysis

the Triton WR-1339 treatment. In some cases, different pro-teins assigned to a given organelle exhibit differences in theirbehavior in a given gradient. For instance, whereas the distri-butions of the peroxisomal membrane protein ABCD3 and theluminal protein urate oxidase (UOX) are similar, they differfrom that of another peroxisomal protein, catalase (CAT), witha greater proportion of the latter sedimenting in the less denseregion of the gradient. The differences in distribution profilesof UOX and CAT were previously noted, leading to the hypoth-esis that a portion of CAT is released from peroxisomes duringcentrifugation through the sucrose gradient (19). Also, whenconsidering the distribution of proteins generally thought to beassociated with ER, the distribution of cytochrome P450CYP2D2, which is associated with the ER membrane with themajority being located on cytosolic surface, differs from that ofthe luminal heat shock proteins HSPA5 and TRA1. Althoughthese differences may be of interest, for this study, the impor-tant finding is that when comparing the control and triton gra-dients, the distribution of the lysosomal proteins can be distin-guished from that of other types of proteins.

The mean distribution profiles for all curated lysosomalproteins are shown in Fig. 4 (Top Panels), with individualspectra plotted for select proteins (Fig. 4, remaining Panels).Note that there is some scatter in the profiles for individualspectra, and this variation needs to be accounted for whenclassifying proteins in the data set (see below). Nonetheless,the mean distribution profiles for all of the curated lysosomalproteins in a given experiment are remarkably similar, withonly one outlier, ACP5, being found in Experiment I (dashedline, Fig. 4 Top Panel). The mean distribution profiles are moretightly clustered in Experiment I than in Experiment II, possiblyreflecting the number of quantifiable spectra analyzed in eachexperiment. Also, there is some animal-to-animal variation inthe degree of density shift induced by Triton WR-1339 treat-ment, and a more pronounced shift was obtained in Experi-ment I. Thus, we chose to use Experiment I for primary clas-sification and Experiment II for corroboration.

Protein Classification—We developed a two-dimensionalindex to facilitate data visualization and analysis. This entailedcalculating the proportion of reporter ion intensity found infractions 1 and 2 of the control and triton gradients (pC1,2 �

[C1 � C2]/[C1�C2�C3�C4] and pT1,2 � [T1 � T2]/[T1�T2�T3�T4], respectively) for each spectrum and thentaking the mean of all spectra assigned to a given protein toprovide a point estimate for its relative enrichment in the lowerdensity fractions. Plotting pT1,2 versus pC1,2 provides a sim-ple way to visualize the degree of the triton shift. Most pro-teins have similar values for pT1,2 and pC1,2, and these “non-shifted” proteins lie on a diagonal in the plots shown in Fig. 5.In contrast, several proteins have a distinct distribution andshow the triton shift, with pT1,2 � pC1,2. Importantly, the lattertype of pattern includes most of the curated lysosomal pro-teins but not the others (black circles and red squares, re-spectively in Fig. 5).

FIG. 4. Distribution profiles for curated lysosomal proteins. TopPanels shows mean distribution profiles for all curated lysosomalproteins. ACP5 is depicted as a dotted black line, all others as solidblack lines. Remaining panels show of individual spectra (black lines)that are used to calculate the mean distribution profiles (thick redline) for classical lysosomal acid phosphatase (ACP2), tartrate-resistant acid phosphatase (ACP5), and cathepsin D (CTSD).

Subcellular Fractionation and Lysosomal Proteomics

10.1074/mcp.M110.006403–6 Molecular & Cellular Proteomics 10.4

Page 7: This paper is available on line at ...cabm-ms.cabm.rutgers.edu/Classification of subcellular location by... · Classification of Subcellular Location by Comparative Proteomic Analysis

We used QDA to estimate the posterior probability that anygiven protein was associated with the lysosome as judged bythe triton shift criteria. For the initial classification set forlysosomal proteins, we used the curated set of lysosomalproteins. For the initial classification set of nonlysosomal pro-teins, we used all other proteins (which may contain somelysosomal proteins). Following QDA, each point is assigned aprobability of lysosomal localization (Supplemental Table S4,Basic.posterior_prob). The black curved dotted line in Fig. 5represents a boundary between the lysosomal and nonlyso-somal distributions (posterior probability � 0.5) for this anal-ysis stage, thus any point falling on this line would be equallylikely to belong to either set.

This initial analysis stage ignores the variability in the dataused to calculate the coordinates pC1,2 and pT1,2. We used aparametric bootstrap procedure, with the parameters derivedfrom the estimated mean and covariance of the measure-ments (see Legend, Fig. 5), to generate 10,000 coordinates foreach protein to simulate its sampling distribution (Fig. 5, grayand pink dots for curated lysosomal and other, respectively).All these coordinates were then used for QDA as above, whichresulted in a new boundary (blue curved dashed line, Fig. 5).Each of the 10,000 points associated with a given proteinwere then assigned a posterior probability and following sort-ing, the values of the 250th and 9750th points used as the95% confidence interval limits (Stage 0 analysis, supple-mental Table S4).

It is possible that proteins in the initial curated lysosomal setwere not true lysosomal residents and that some proteins not

on the curated list are actually lysosomal residents and wereinitially classified as nonlysosomal. To address this, proteinswere reassigned to the lysosomal classification set if the lower95% confidence limit was � 0.5, reassigned to the nonlyso-somal classification set if the upper 95% confidence limitwas � 0.5, or reassigned to a third “ambiguous” set if they didnot meet either criterion. We then used the 10,000 pointsassigned to each protein in the first two sets (excluding am-biguous proteins) in the QDA procedure to recalculate poste-rior probability estimates and confidence intervals for eachprotein as above (Stage 1 analysis, supplemental Table S4).We again reassigned proteins as before to either the lyso-somal, nonlysosomal, or ambiguous sets and the QDA pro-cedure repeated (Stage 2 analysis, supplemental Table S4).All assignments stabilized following the second and first iter-ation for Experiments 1 and 2, respectively, and the finalboundary is shown as a solid green line in Fig. 5.

Fig. 6 shows the distribution of proteins identified in Exper-iment I based on their Stage 2 classifications, with corre-sponding data for Experiment II in supplemental Fig. S4.These classifications should be considered tentative (see Dis-cussion), but provide a useful starting point for further inves-tigation. Of the proteins in the curated lysosomal set, all butone (ACP5, tartrate-resistant acid phosphatase) was eitherclassified as having lysosomal or ambiguous distributions(Fig. 6, Top Left Panel). Of the proteins not in the curated set,five were classified as lysosomal, 15 as ambiguous, and theremaining 487 as nonlysosomal (Fig. 6, Top Right Panel). Ourdata set overlapped with that of two prior lysosomal proteo-

FIG. 5. Classification of proteins using quadratic discriminant analysis (QDA). Each protein identified in a given experiment isrepresented by a separate symbol: black circles, curated lysosomal proteins; red squares, others. The gray and red dots depict the bootstrappoints of the curated lysosomal and other data set, respectively. Curves represent the boundary lines where the posterior probability forassignment as lysosomal or not lysosomal is equal for different classification routines (see text for details). Statistical analysis was conductedas follows: For each protein we generated 10,000 points using a bivariate normal parametric bootstrap procedure using the two sample meansof pC1,2 and pT1,2 (calculated from individual spectra associated with each protein) and their variances and the covariance. Formally, denoteby pC and pT the sample means of the n reporter ion intensity measurements for a particular protein and let sC

2 , sT2, and sCT denote, respectively,

the variances of the n spectra and their covariance. Then the precision of the estimates of pC and pT are given by sC2/n, sT

2/n, and sCT/n. Thesethree parameters, which reflect the precision of the estimates, are used to generate the aforementioned bivariate normal random variables. (Forthose proteins with fewer than three spectra, we used estimates of sC

2 , sT2, and sCT derived from the average values of these parameters across

all proteins.) This process was carried out for each of the proteins (resulting in 5,440,000 and 4,500,000 points for Experiments 1 and 2,respectively), and these points were used to carry out the discriminant analysis. This procedure effectively places greater weight on proteinswhich are more accurately characterized, i.e. those with smaller variances and covariances.

Subcellular Fractionation and Lysosomal Proteomics

Molecular & Cellular Proteomics 10.4 10.1074/mcp.M110.006403–7

Page 8: This paper is available on line at ...cabm-ms.cabm.rutgers.edu/Classification of subcellular location by... · Classification of Subcellular Location by Comparative Proteomic Analysis

mics studies. One study reported identification of 215 pro-teins in a rat liver tritosome integral membrane preparation(27). We identified 88 of these and classified 11 as lysosomal,9 as ambiguous, and 68 as nonlysosomal (Fig. 6, Middle LeftPanel). Another study classified 145 human placental proteinsas being associated with the lysosomal membrane (28). Weidentified 24 rat orthologs of these and classified 7 as lyso-somal, 7 as ambiguous, and 10 as nonlysosomal (Fig. 6,Middle Right Panel). We identified 84 proteins that were listedin the MitoCarta database of mitochondrial proteins (29) and

all were classified here as nonlysosomal (Fig. 6, Bottom LeftPanel), whereas of the 30 proteins identified that were alsolisted in the Peroxisome Database (30), one was classified aslysosomal and the remaining as nonlysosomal (Fig. 6, BottomRight Panel).

Subcellular Localization of Selected Candidates—Fig. 7shows the confidence intervals for the Stage II distributions ofall 37 curated lysosomal proteins identified in Experiment I aswell as an equal number of proteins in the curated othercategory that had the highest posterior probability scores.Corresponding data for Experiment II are in Supplemen-tal Fig. S5. We chose to investigate the distribution of twoproteins where good antibodies were available for quantitativeWestern blotting: Cu/Zn superoxide dismutase 1 (SOD1),which had a clear lysosomal distribution in both experiments,and ATP-binding cassette subfamily B (MDR/TAP) member 6(ABCB6), which fell into the ambiguous category.

Classical differential centrifugation analysis was used tocompare the distribution of SOD1 and ABCB6 to that ofestablished markers for lysosomes (�-galactosidase), peroxi-somes (catalase), mitochondria (cytochrome oxidase), andendoplasmic reticulum (glucose 6-phosphatase) (Fig. 8).ABCB6 exhibited a pattern consistent with either a lyso-somal or peroxisomal distribution, with the greatest enrich-ment in the L fraction and the majority of the total present in

FIG. 6. Classifications of different data sets. Selected proteinsidentified in Experiment I are represented according to their Stage 2assignments: Black circles, lysosomal; red squares, nonlysosomal;gray diamonds, ambiguous. Dashed line represents points with apredicted posterior probability of 0.5. Top Left Panel, contained in theinitial curated lysosomal set. Top Right Panel, not in the initial curatedlysosomal set. Middle Left Panel, listed in Table I of Bagshaw et al.(27) Middle Right Panel, listed in Table I of Schroeder et al. (28),Bottom Panels. Bottom Left Panel, proteins assigned to mouse mi-tochondria with confidence levels of 1–5 (http://www.broadinstitute-.org/pubs/MitoCarta/index.html) (29). Bottom Right Panel: listed in therat and/or mouse pages of the PeroxisomeDB 2.0 (http://www.peroxi-somedb.org/) (30).

FIG. 7. Posterior probabilities of select proteins for assignmentas lysosomal. Symbols are as in Fig. 6. Error bars indicate 95%confidence intervals. For clarity of presentation, the mouse gene nameis used when the rat and mouse ortholog differ. Mouse/rat substitutionsare as follows: NPC1/B2RYH4_RAT, LAMP1/LAMP1_RAT, GNS/Q32KJ5_RAT, HEXB/HEXB_RAT, GUSB/BGLR_RAT.

Subcellular Fractionation and Lysosomal Proteomics

10.1074/mcp.M110.006403–8 Molecular & Cellular Proteomics 10.4

Page 9: This paper is available on line at ...cabm-ms.cabm.rutgers.edu/Classification of subcellular location by... · Classification of Subcellular Location by Comparative Proteomic Analysis

the M and L fractions together. The bulk of cellular SOD1was detected in the S fraction, which is consistent with theestablished residence of this protein in the cytoplasm. How-ever, there was significant enrichment of SOD1 within the Lfraction, consistent with a secondary association with lyso-somes or peroxisomes.

Fig. 9 shows sucrose density gradients of the ML fractionwhich together contains the majority of the mitochondrial,peroxisomal, and lysosomal markers. The distribution ofABCB6 followed that of the lysosomal marker �-galactosid-ase in both the control and triton gradients. SOD1 had a morecomplex distribution. For the control rat liver sample, a portionof the SOD1 remained at the top of the gradient, as might beexpected for a cytosolic protein contaminant of a differentialcentrifugation fraction, whereas the remainder had a distribu-tion similar to that of the lysosomal marker. SOD1 was clearlyshifted in the Triton WR-1339 treated sample, again consis-tent with a portion of the protein being associated withlysosomes.

DISCUSSION

Several previous studies have combined subcellular frac-tionation with mass spectrometry and protein sequencing inthe proteomic analysis of lysosomes and related organelles(27, 28, 31–34). These have provided valuable insights into themembrane and lumenal composition of the lysosome butinterpretation is complicated by the presence of other organ-elles in the preparations. In one of these studies, mass spec-trometry was conducted at different stages in the purificationof lysosomal membranes and spectral counting used to helpdistinguish likely lysosomal proteins, which are increasinglyenriched during purification, from contaminants, which areincreasingly depleted (28). Here, a lysosome-enriched prepa-ration was subjected to an in vitro treatment that preferentiallyruptures lysosomes and the final purification step involvedenrichment of the lysed membranes. This approach doesrepresent a significant step forward in terms of distinguishinglysosomal proteins from contaminants but it is limited to lyso-somal membrane proteins.

In this study, we have taken an alternative approach thatcan be used to classify both luminal and membrane proteinsas lysosomal or nonlysosomal. Here, we use an isobaric la-beling strategy to quantitate the shift in density of the lyso-some that is induced by treatment with Triton WR-1339 anduse a set of curated lysosomal proteins to establish criteria toguide the assignment of other proteins to this organelle. It isworth considering inherent assumptions and limitations in thecurrent experimental design. First, in terms of mass spec-trometry, there are several caveats in protein quantificationusing isobaric labels, including potential artifacts arising fromanalysis of mixed spectra (35). We have attempted to accountfor variable data quality by averaging spectra and includingerror estimates, but this remains a concern, especially forproteins with sparse coverage. Second, in terms of bioinfor-

FIG. 8. Distribution of lysosomal candidates in control and tri-ton-treated rat liver differential centrifugation fractions. For eachplot, area is proportional to total signal. Left Panels, Controls, RightPanels, Triton WR-1339 treated. Ordinate, relative specific activity (per-centage of total recovered activity or signal normalized to percentage oftotal recovered protein). Abscissa, relative protein content of fraction(cumulative from left to right). Fractions are: N, nuclear; M, heavy mito-chondrial; L, light mitochondrial; P, microsomal and; S, high speedsupernatant. Markers were measured by activity assays whereas ABC6and SOD1 were measured by quantitative Western blotting analyzingequal amounts of protein (2 and 4 �g) for each fraction.

Subcellular Fractionation and Lysosomal Proteomics

Molecular & Cellular Proteomics 10.4 10.1074/mcp.M110.006403–9

Page 10: This paper is available on line at ...cabm-ms.cabm.rutgers.edu/Classification of subcellular location by... · Classification of Subcellular Location by Comparative Proteomic Analysis

matics, it should be noted that there are conflicting reportsregarding cellular localization of some proteins, and the com-position of the curated lysosomal set is somewhat subjective.However, the iterative method used to obtain the Stage IIclassifications should tolerate a certain level of misclassifica-tion. Finally, in terms of subcellular fractionation, it should benoted that our initial analysis was conducted on L differentialcentrifugation fractions that only represent a portion of the

lysosomes (15–20% based on marker enzyme analysis) andprotein (�2–2.5%) present in the liver homogenate. Based onthe postulate of biochemical homogeneity, lysosomes in the Lfraction are likely to have a composition that is representativeof the entire sample (9). However, one cannot assume that thepostulate of single location, which is obeyed to a first approx-imation by well-established organelle markers, holds true forlysosomal candidates. It is also possible that some peripheralproteins associated with the cytoplasmic leaflet of the lyso-somal membrane dissociate during the fractionation proce-dure, resulting in a false negative classification. Thus, ourclassifications should be considered as a guide for furtherinvestigation using balance sheet and other types of analyses.Select proteins of interest that were not in the initial curatedlysosomal data set and had posterior probability scoreshigher than that of ABCB6 in Experiment I are discussedbelow in order of their score.

Phospholipase B Domain Containing 1 (PLBD1)—PLBD1(also known as LAMA-like protein 1 and FLJ22662) was clas-sified as lysosomal in Experiments I and II. This protein is aparalog of a newly discovered lysosomal protein, phospho-lipase B domain containing 2 (also named P76, LAMA-likeprotein 2, and LOC196463) (15, 17). PLBD1 was previouslyfound in preparations of Man6-P glycoproteins from varioussources (1, 23) and was directly shown to contain Man6-Presidues (36). These findings and the results presented hereare together highly suggestive that PLBD1 is a bona fidelysosomal protein.

Cu/Zn Superoxide Dismutase 1—SOD1 catalyzes the con-version of superoxide anions into oxygen and hydrogen per-oxide. In our initial analysis, SOD1 was classified as lysosomalin both Experiments I and II. We subsequently performedWestern blotting analysis of differential centrifugation frac-tions and ML sucrose density gradient fractions from controland triton treated rats (Figs. 8 and 9). This indicated thatSOD1 is associated with the lysosome, but that this organelleis not its primary residence. Although SOD1 is generally re-garded as a cytosolic protein, some reports have shown thatit can be found in the mitochondrial intermembrane space (37,38). A recent report has identified SOD1 in a peroxisome-enriched subcellular fraction and, based on immunofluores-cence microscopy studies on cells overexpressing bothSOD1 and “copper chaperone of SOD1,” concluded that cop-per chaperone of SOD1 mediates import of SOD1 into per-oxisomes (39).

Although SOD1 clearly has a complex distribution, it isimportant to determine whether the minor amount present inmembrane fractions under physiological conditions is asso-ciated with mitochondria, peroxisomes, or lysosomes. A rig-orous earlier study reported that small amounts of SOD1(ranging from 2 to 8% of the total, higher under starvationconditions) are associated with the lysosome, likely reflect-ing autophagy of cytosolic protein (40). This was corrobo-rated by a study employing cryosection immunoelectron

FIG. 9. Distribution of ABCB6 and SOD1 in sucrose densitygradients of control and triton-treated rat liver ML fractions. Redand black symbols represent analyses of samples from Triton-WR1339-treated and control animals, respectively. The distribution ofABCB6 and SOD1 was determined by quantitative Western blottinganalyzing equivalent volume proportions (corresponding to 0.5 and/or1 mg wet weight liver) for each fraction. Western blotting for SOD1revealed a single band that migrated with electrophoretic mobilitybetween the 11 and 31 kDa size markers (data not shown), which isconsistent with the known mass of the rat protein (�20 kDa). Westernblotting for ABCB6 revealed several bands but the one shown whichis shifted in response to Triton-treatment was the only protein migrat-ing between the 59 and 110 kDa size standards (data not shown),consistent with the observed size of rat ABCB6 (�80 kDa).

Subcellular Fractionation and Lysosomal Proteomics

10.1074/mcp.M110.006403–10 Molecular & Cellular Proteomics 10.4

Page 11: This paper is available on line at ...cabm-ms.cabm.rutgers.edu/Classification of subcellular location by... · Classification of Subcellular Location by Comparative Proteomic Analysis

microscopy, which also showed that SOD1 is relativelyresistant to lysosomal proteolysis (41). The physiologicalrationale for the presence of SOD1 within the lysosomeremains to be elucidated. It is possible that it is targeted tothe lysosome for degradation but relatively stable within thisenvironment and thus hydrolyzed slowly. Alternatively, itmay be relevant that the activity of SOD1 is relatively inde-pendent of pH (42), which may allow it to function within thelysosome.

Glutamyl Aminopeptidase—Glutamyl aminopeptidase wasclassified as lysosomal in Experiment I and ambiguous inExperiment II. It appears to play a role in the control of bloodpressure via degradation of angiotensin II (43) and our resultssuggest a lysosomal localization for this protein. Glutamylaminopeptidase has also been found to be significantly en-riched in lysosomal membranes isolated from human placenta(28).

Vacuolar ATPase V0 Domain Subunit d1 and other Compo-nents of the Vacuolar ATPases—Although vacuolar ATPasesare responsible for lysosomal acidification, these protonpumps are localized to a variety of membranes and are com-posed of multiple subunits and isoforms (44). The subunitsalso undergo dynamic association and dissociation. Becauseof this complexity, we did not place any of the subunits in ourinitial list of curated lysosomal proteins. Nonetheless, anumber were identified, with the following classifications:ATP6V0D1, lysosomal in Experiments I and II; TCIG1(ATP6V0A3) and ATP6AP1, ambiguous in both experiments;ATP6V0A1, ambiguous in Experiment I and not found in Ex-periment II; ATP6V1A1, ambiguous in Experiment II, nonlyso-somal in Experiment I; and ATP6V1B2, not lysosomal in Ex-periments I and II. Although the nonlysosomal classification ofsome of the subunits of the V1 domain may reflect subcellularlocalization, these are peripheral proteins, and it is possiblethat they may dissociate from membranes during the fraction-ation process.

ADP-ribosylation Factor-like Protein 8B—ARL8B was clas-sified as lysosomal in Experiment I and ambiguous in Exper-iment II. It belongs to the Arf-like family of small GTP-ases andwas reported to participate in chromosome segregation andto localize with microtubules on the mitotic spindle (45).ARLB8 was identified in previous lysosomal proteomics stud-ies (27, 28) and a lysosomal localization is consistent withvisualization of fusion protein chimeras and effects of overex-pression on lysosome distribution within cells (46, 47).

Amyloid P-component, Serum (APCS)—APCS was classi-fied as ambiguous in Experiment I and lysosomal in Experi-ment II. This protein is present in plasma and cerebrospinalfluid and may be involved in the pathogenesis of Alzheimerdisease as it appears to prevent proteolysis of amyloid fibrils(48, 49). It was previously found in preparations of Man6-Pglycoproteins from human and mouse plasma (50, 51) as wellas a wide range of rat tissues (23) although the presence ofMan6-P has not been directly demonstrated. Interestingly,

there is evidence that APCS can bind carbohydrates andproteins containing Man6-P (52) which raises the possibilitythat this protein may not be a Man6-P glycoprotein per se butmay be transported to the lysosome by MPR-mediated en-docytosis while in association with other proteins containingMan6-P.

Ferric-chelate Reductase 1—Ferric-chelate reductase (alsocalled for stromal cell-derived receptor 2) was classified asambiguous in Experiment I and lysosomal in Experiment II.This protein belongs to the cytochrome b 561 family and is aferric-reductase (53). LCTYB, another member of the Cyto-chrome b 561 family with ferric-reductase activity, has beenreported to localize to the late-endosome/lysosomal mem-brane (54).

ATP-binding Cassette Subfamily B {MDR/TAP} Member6—ABCB6 had an ambiguous localization as determined inboth Experiment I and II, with several peptides clearly exhib-iting the triton shift. This protein is an ATP binding cassettetransporter, being part of a large family of proteins that playimportant roles in the transport of a variety of substratesacross membranes (55). There is considerable disagreementregarding the intracellular distribution of ABCB6. Initially de-scribed as a mitochondrial protein involved in iron homeosta-sis (56), ABCB6 subsequently was reported to reside in themitochondrial outer membrane (57), in both the mitochondrialouter membrane and the plasma membrane (58), in endoplas-mic reticulum derived compartments consisting mainly ofGolgi (59), and in late endosomes/lysosomes (60). In previousproteomics studies, ABCB6 was identified in rat tritosomes(27) and was enriched in lysosomal membranes from humanplacenta (28). A variety of morphological and/or subcellularfractionation approaches were used in these localization stud-ies, but it is worth noting that in some of them, the presenceof lysosomes in preparations of “purified” mitochondria wasnot fully appreciated. Our balance sheet analysis providesstrong support for the study reporting a lysosomal residencefor ABCB6 (60).

Other—It is also possible that additional proteins classi-fied as either ambiguous or nonlysosomal from this analysiscontribute to lysosomal function, and thus represent falsenegatives. For instance, of the curated lysosomal proteins,tartrate-resistant acid phosphatase (ACP5) was classifiedas nonlysosomal in both Experiments I and II. This mayreflect experimental error associated with measurements onrelatively few spectra (three in Experiment I, one in Experi-ment II). Alternatively, it is possible that the data reflectunderlying biology. ACP5 is expressed in high levels inosteoclasts where it is involved in bone resorption (61) butis also present in numerous other cell types where it func-tions in dephosphorylation of Man 6-P containing lysosomalproteins (62). Interestingly, early studies suggested a com-plex compartmentalization of the dephosphorylation activity(63). Additional investigation will be needed to resolve thisissue.

Subcellular Fractionation and Lysosomal Proteomics

Molecular & Cellular Proteomics 10.4 10.1074/mcp.M110.006403–11

Page 12: This paper is available on line at ...cabm-ms.cabm.rutgers.edu/Classification of subcellular location by... · Classification of Subcellular Location by Comparative Proteomic Analysis

Future prospects. The current study was initiated as astep toward the goal of simultaneously identifying and clas-sifying proteins with respect to lysosomal residence. Al-though providing useful information, there are several direc-tions in which this approach can be optimized. First, wehave found that the MALDI-based workflow provided rea-sonable quantitative data but coverage was limited, identi-fying only around half of the soluble lysosomal proteinspreviously detected in a rat liver Man 6-P glycoprotein prep-aration (23). This problem will likely be exacerbated in thecase of the less abundant components of the lysosome,which may include “new” lysosomal proteins that are ofconsiderable interest. However, we anticipate that the nextgeneration of experiments using LC-MS/MS based methodsfor quantitative mass spectrometry combined with multi-dimensional fractionation of the sample should address thisproblem. Second, it may be worthwhile to extend this ap-proach from liver to other tissues, most of which are refrac-tory to the Triton WR-1339 treatment. The density shiftcould be accomplished using mutant mice deficient in thelysosomal lipid transport protein NPC2, which exhibit alysosomal accumulation of cholesterol and other lipids anda characteristic density shift (64), or by treating cells withagents to induce a lipidosis (65). Finally, when interpretingthe biological relevance of a lysosomal localization, it isessential to determine the total cellular distribution of a candi-date of interest. For example, if most or all of the cellular pop-ulation of a given protein is located within the lysosome, thiswould argue strongly for a biological function within this organ-elle. This is not the case if only a small proportion is found to belysosomal, a situation which may instead suggest that this pro-tein is mistargeted to the lysosome or is targeted there forcatabolic purposes. This may be the case with SOD1, whereonly a small proportion of the total cellular SOD1 appears to belysosomal. In this study, we have determined the total cellulardistribution of select candidates of interest by quantitativeWestern blotting but the application of quantitative mass spec-trometry to measure protein levels in differential centrifugationfractions should make this achievable in a more efficient andglobal manner.

Acknowledgments—We thank A. Dautreloux, J. Thirion and I. So-har for help with subcellular fractionation and enzyme assays.

* This work was supported by grants from the National Institutes ofHealth (R01DK054317 and S10RR020932) (P. L.) and FRFC grant 24543 08 (M. J.).

□S This article contains supplemental Tables S1 to S4 andsupplemental Figs. S1 to S5.

** To whom correspondence should be addressed: CABM, 679Hoes Lane, Piscataway, NJ 08854, USA. Tel.: 732-235-5032; Fax:732-235-4466; E-mail: [email protected].

REFERENCES

1. Lubke, T., Lobel, P., and Sleat, D. E. (2009) Proteomics of the lysosome.Biochim. Biophys. Acta 1793, 625–635

2. Journet, A., and Ferro, M. (2004) The potentials of MS-based subproteomic

approaches in medical science: the case of lysosomes and breast can-cer. Mass Spectrom. Rev. 23, 393–442

3. Lee, J. H., Yu, W. H., Kumar, A., Lee, S., Mohan, P. S., Peterhoff, C. M.,Wolfe, D. M., Martinez-Vicente, M., Massey, A. C., Sovak, G., Uchiyama,Y., Westaway, D., Cuervo, A. M., and Nixon, R. A. (2010) Lysosomalproteolysis and autophagy require presenilin 1 and are disrupted byAlzheimer-related PS1 mutations. Cell 141, 1146–1158

4. Sidransky, E., Nalls, M. A., Aasly, J. O., Aharon-Peretz, J., Annesi, G.,Barbosa, E. R., Bar-Shira, A., Berg, D., Bras, J., Brice, A., Chen, C. M.,Clark, L. N., Condroyer, C., De, Marco, E. V., Durr, A., Eblan, M. J.,Fahn, S., Farrer, M. J., Fung, H. C., Gan-Or, Z., Gasser, T., Gershoni-Baruch, R., Giladi, N., Griffith, A., Gurevich, T., Januario, C., Kropp, P.,Lang, A. E., Lee-Chen, G. J., Lesage, S., Marder, K., Mata, I. F.,Mirelman, A., Mitsui, J., Mizuta, I., Nicoletti, G., Oliveira, C., Ottman,R., Orr-Urtreger, A., Pereira, L. V., Quattrone, A., Rogaeva, E., Rolfs,A., Rosenbaum, H., Rozenberg, R., Samii, A., Samaddar, T., Schulte,C., Sharma, M., Singleton, A., Spitz, M., Tan, E. K., Tayebi, N., Toda,T., Troiano, A. R., Tsuji, S., Wittstock, M., Wolfsberg, T. G., Wu, Y. R.,Zabetian, C. P., Zhao, Y., and Ziegler, S. G. (2009) Multicenter analysisof glucocerebrosidase mutations in Parkinson’s disease. N. Engl.J. Med. 361, 1651–1661

5. Sun, B., Zhou, Y., Halabisky, B., Lo, I., Cho, S. H., Mueller-Steiner, S.,Devidze, N., Wang, X., Grubb, A., and Gan, L. (2008) Cystatin C-cathe-psin B axis regulates amyloid beta levels and associated neuronal defi-cits in an animal model of Alzheimer’s disease. Neuron 60, 247–257

6. Callahan, J. W., Bagshaw, R. D., and Mahuran, D. J. (2009) The integralmembrane of lysosomes: its proteins and their roles in disease. J. Pro-teomics 72, 23–33

7. Sleat, D. E., Jadot, M., and Lobel, P. (2007) Lysosomal proteomics anddisease. Proteomics - Clinical Applications 1, 1134–1146

8. Schroder, B. A., Wrocklage, C., Hasilik, A., and Saftig, P. (2010) Theproteome of lysosomes. Proteomics 10, 4053–4076

9. de Duve, C. (1971) Tissue fractionation. Past and present. J. Cell Biol. 50,20d–55d

10. de Duve, C., Pressman, B. C., Gianetto, R., Wattiaux, R., and Appelmans,F. (1955) Tissue fractionation studies. 6. Intracellular distribution patternsof enzymes in rat-liver tissue. Biochem. J. 60, 604–617

11. Wattiaux, R., Wibo, M., and Baudhuin, P. (1963) Influence of the injection ofTriton WR-1339 on the properties of rat-liver lysosomes. In: de Reuck,A. V. S., and Cameron, M. P., eds. Ciba Foundation Symposium Lyso-somes, pp. 176–200, Little, Brown, and Company, Boston

12. Sohar, I., Sleat, D. E., Jadot, M., and Lobel, P. (1999) Biochemical charac-terization of a lysosomal protease deficient in classical late infantileneuronal ceroid lipofuscinosis (LINCL) and development of an enzyme-based assay for diagnosis and exclusion of LINCL in human specimensand animal models. J. Neurochem. 73, 700–711

13. Naureckiene, S., Sleat, D. E., Lackland, H., Fensom, A., Vanier, M. T.,Wattiaux, R., Jadot, M., and Lobel, P. (2000) Identification of HE1 as theSecond Gene of Niemann-Pick C Disease. Science 290, 2298–2301

14. Boonen, M., Hamer, I., Boussac, M., Delsaute, A. F., Flamion, B., Garin, J.,and Jadot, M. (2006) Intracellular localization of p40, a protein identifiedin a preparation of lysosomal membranes. Biochem. J. 395, 39–47

15. Deuschl, F., Kollmann, K., von, Figura, K., and Lubke, T. (2006) Molecularcharacterization of the hypothetical 66.3-kDa protein in mouse: lyso-somal targeting, glycosylation, processing and tissue distribution. FEBSLett. 580, 5747–5752

16. Schieweck, O., Damme, M., Schroder, B., Hasilik, A., Schmidt, B., andLubke, T. (2009) NCU-G1 is a highly glycosylated integral membraneprotein of the lysosome. Biochem. J. 422, 83–90

17. Jensen, A. G., Chemali, M., Chapel, A., Kieffer-Jaquinod, S., Jadot, M.,Garin, J., and Journet, A. (2007) Biochemical characterization and lyso-somal localization of the mannose-6-phosphate protein p76 (hypotheti-cal protein LOC196463). Biochem. J. 402, 449–458

18. Leighton, F., Poole, B., Beaufay, H., Baudhuin, P., Coffey, J. W., Fowler,S., and de Duve, C. (1968) The large-scale separation of peroxisomes,mitochondria, and lysosomes from the livers of rats injected with tritonWR-1339. Improved isolation procedures, automated analysis, bio-chemical and morphological properties of fractions. J. Cell Biol. 37,482–513

19. Beaufay, H., Jacques, P., Baudhuin, P., Sellinger, O. Z., Berthet, J., and deDuve, C. (1964) Tissue fractionation studies. 18. Resolution of mitochon-

Subcellular Fractionation and Lysosomal Proteomics

10.1074/mcp.M110.006403–12 Molecular & Cellular Proteomics 10.4

Page 13: This paper is available on line at ...cabm-ms.cabm.rutgers.edu/Classification of subcellular location by... · Classification of Subcellular Location by Comparative Proteomic Analysis

drial fractions from rat liver into three distinct populations of cytoplasmicparticles by means of density equilibration in various gradients. Biochem.J. 92, 184–205

20. Sleat, D. E., Sohar, I., Lackland, H., Majercak, J., and Lobel, P. (1996) Ratbrain contains high levels of mannose-6-phosphorylated glycoproteinsincluding lysosomal enzymes and palmitoyl-protein thioesterase, an en-zyme implicated in infantile neuronal lipofuscinosis. J. Biol. Chem. 271,19191–19198

21. Storrie, B., and Madden, E. A. (1990) Isolation of subcellular organelles.Methods Enzymol. 182, 203–225

22. Aebi, H. (1984) Catalase in vitro. Methods Enzymol. 105, 121–12623. Sleat, D. E., Della Valle, M. C., Zheng, H., Moore, D. F., and Lobel, P. (2008)

The Mannose 6-Phosphate Glycoprotein Proteome. J. Proteome Res. 7,3010–3021

24. Choe, L., D’Ascenzo, M., Relkin, N. R., Pappin, D., Ross, P., Williamson, B.,Guertin, S., Pribil, P., and Lee, K. H. (2007) 8-plex quantitation ofchanges in cerebrospinal fluid protein expression in subjects undergoingintravenous immunoglobulin treatment for Alzheimer’s disease. Pro-teomics 7, 3651–3660

25. Craig, R., and Beavis, R. C. (2004) TANDEM: matching proteins with tan-dem mass spectra. Bioinformatics 20, 1466–1467

26. Venables, W. N., and Ripley, B. D. (2002) Modern Applied Statistics with S,4 Ed., Springer, New York

27. Bagshaw, R. D., Mahuran, D. J., and Callahan, J. W. (2005) A proteomicanalysis of lysosomal integral membrane proteins reveals the diversecomposition of the organelle. Mol. Cell Proteomics 4, 133–143

28. Schroder, B., Wrocklage, C., Pan, C., Jager, R., Kosters, B., Schafer, H.,Elsasser, H. P., Mann, M., and Hasilik, A. (2007) Integral and associatedlysosomal membrane proteins. Traffic 8, 1676–1686

29. Pagliarini, D. J., Calvo, S. E., Chang, B., Sheth, S. A., Vafai, S. B., Ong, S. E.,Walford, G. A., Sugiana, C., Boneh, A., Chen, W. K., Hill, D. E., Vidal, M.,Evans, J. G., Thorburn, D. R., Carr, S. A., and Mootha, V. K. (2008) Amitochondrial protein compendium elucidates complex I disease biol-ogy. Cell 134, 112–123

30. Schluter, A., Real-Chicharro, A., Gabaldon, T., Sanchez-Jimenez, F., andPujol, A. (2010) PeroxisomeDB 2.0: an integrative view of the globalperoxisomal metabolome. Nucleic Acids Res. 38, D800–805

31. Chataway, T. K., Whittle, A. M., Lewis, M. D., Bindloss, C. A., Davey, R. C.,Moritz, R. L., Simpson, R. J., Hopwood, J. J., and Meikle, P. J. (1998)Two-dimensional mapping and microsequencing of lysosomal proteinsfrom human placenta. Placenta 19, 643–654

32. Chi, A., Valencia, J. C., Hu, Z. Z., Watabe, H., Yamaguchi, H., Mangini, N. J.,Huang, H., Canfield, V. A., Cheng, K. C., Yang, F., Abe, R., Yamagishi, S.,Shabanowitz, J., Hearing, V. J., Wu, C., Appella, E., and Hunt, D. F.(2006) Proteomic and bioinformatic characterization of the biogenesisand function of melanosomes. J. Proteome Res. 5, 3135–3144

33. Casey, T. M., Meade, J. L., and Hewitt, E. W. (2007) Organelle proteomics:identification of the exocytic machinery associated with the natural killercell secretory lysosome. Mol. Cell Proteomics 6, 767–780

34. Tribl, F., Gerlach, M., Marcus, K., Asan, E., Tatschner, T., Arzberger, T.,Meyer, H. E., Bringmann, G., and Riederer, P. (2005) ”Subcellular pro-teomics“ of neuromelanin granules isolated from the human brain. Mol.Cell Proteomics 4, 945–957

35. Ow, S. Y., Salim, M., Noirel, J., Evans, C., Rehman, I., and Wright, P. C.(2009) iTRAQ underestimation in simple and complex mixtures: “thegood, the bad and the ugly”. J. Proteome Res. 8, 5347–5355

36. Sleat, D. E., Zheng, H., Qian, M., and Lobel, P. (2006) Identification of sitesof mannose 6-phosphorylation on lysosomal proteins. Mol. Cell Pro-teomics 5, 686–701

37. Okado-Matsumoto, A., and Fridovich, I. (2001) Subcellular Distribution ofSuperoxide Dismutases (SOD) in Rat Liver. Cu, Zn-SOD in mitochondria.J. Biol. Chem. 276, 38388–38393

38. Kawamata, H., and Manfredi, G. (2008) Different regulation of wild-type andmutant Cu, Zn superoxide dismutase localization in mammalian mito-chondria. Hum. Mol. Genet. 17, 3303–3317

39. Islinger, M., Li, K. W., Seitz, J., Volkl, A., and Luers, G. H. (2009) Hitchhikingof Cu/Zn superoxide dismutase to peroxisomes–evidence for a naturalpiggyback import mechanism in mammals. Traffic 10, 1711–1721

40. Geller, B. L., and Winge, D. R. (1982) Rat liver Cu, Zn-superoxide dismu-tase. Subcellular location in lysosomes. J. Biol. Chem. 257, 8945–8952

41. Rabouille, C., Strous, G. J., Crapo, J. D., Geuze, H. J., and Slot, J. W. (1993)

The differential degradation of two cytosolic proteins as a tool to monitorautophagy in hepatocytes by immunocytochemistry. J. Cell Biol. 120,897–908

42. Ellerby, L. M., Cabelli, D. E., Graden, J. A., and Valentine, J. S. (1996)Copper-zinc superoxide dismutase: Why not pH-dependent? J. Am.Chem. Soc. 118, 6556–6561

43. Harata, T., Ando, H., Iwase, A., Nagasaka, T., Mizutani, S., and Kikkawa, F.(2006) Localization of angiotensin II, the AT1 receptor, angiotensin-converting enzyme, aminopeptidase A, adipocyte-derived leucine amino-peptidase, and vascular endothelial growth factor in the human ovarythroughout the menstrual cycle. Fertil. Steril. 86, 433–439

44. Toei, M., Saum, R., and Forgac, M. (2010) Regulation and isoform functionof the V-ATPases. Biochemistry 49, 4715–4723

45. Okai, T., Araki, Y., Tada, M., Tateno, T., Kontani, K., and Katada, T. (2004)Novel small GTPase subfamily capable of associating with tubulin isrequired for chromosome segregation. J. Cell Sci. 117, 4705–4715

46. Bagshaw, R. D., Callahan, J. W., and Mahuran, D. J. (2006) The Arf-familyprotein, Arl8b, is involved in the spatial distribution of lysosomes.Biochem. Biophys. Res. Commun. 344, 1186–1191

47. Hofmann, I., and Munro, S. (2006) An N-terminally acetylated Arf-likeGTPase is localised to lysosomes and affects their motility. J. Cell Sci.119, 1494–1503

48. Kolstoe, S. E., Ridha, B. H., Bellotti, V., Wang, N., Robinson, C. V., Crutch,S. J., Keir, G., Kukkastenvehmas, R., Gallimore, J. R., Hutchinson, W. L.,Hawkins, P. N., Wood, S. P., Rossor, M. N., and Pepys, M. B. (2009)Molecular dissection of Alzheimer’s disease neuropathology by deple-tion of serum amyloid P component. Proc. Natl. Acad. Sci. U.S.A. 106,7619–7623

49. Tennent, G. A., Lovat, L. B., and Pepys, M. B. (1995) Serum amyloid Pcomponent prevents proteolysis of the amyloid fibrils of Alzheimer dis-ease and systemic amyloidosis. Proc. Natl. Acad. Sci. U.S.A. 92,4299–4303

50. Sleat, D. E., Wang, Y., Sohar, I., Lackland, H., Li, Y., Li, H., Zheng, H., andLobel, P. (2006) Identification and Validation of Mannose 6-PhosphateGlycoproteins in Human Plasma Reveal a Wide Range of Lysosomal andNon-lysosomal Proteins. Mol. Cell Proteomics 5, 1942–1956

51. Qian, M., Sleat, D. E., Zheng, H., Moore, D., and Lobel, P. (2008) Proteo-mics Analysis of Serum from Mutant Mice Reveals Lysosomal ProteinsSelectively Transported by Each of the Two Mannose 6-PhosphateReceptors. Mol. Cell Proteomics 7, 58–70

52. Loveless, R. W., Floyd-O’Sullivan, G., Raynes, J. G., Yuen, C. T., and Feizi,T. (1992) Human serum amyloid P is a multispecific adhesive proteinwhose ligands include 6-phosphorylated mannose and the 3-sulphatedsaccharides galactose, N-acetylgalactosamine and glucuronic acid.EMBO J. 11, 813–819

53. Vargas, J. D., Herpers, B., McKie, A. T., Gledhill, S., McDonnell, J., van denHeuvel, M., Davies, K. E., and Ponting, C. P. (2003) Stromal cell-derivedreceptor 2 and cytochrome b561 are functional ferric reductases.Biochim. Biophys. Acta 1651, 116–123

54. Zhang, D. L., Su, D., Berczi, A., Vargas, A., and Asard, H. (2006) Anascorbate-reducible cytochrome b561 is localized in macrophage lyso-somes. Biochim. Biophys. Acta 1760, 1903–1913

55. Holland, K. A., and Holland, I. B. (2005) Adventures with ABC-proteins:highly conserved ATP-dependent transporters. Acta Microbiol. Immunol.Hung. 52, 309–322

56. Mitsuhashi, N., Miki, T., Senbongi, H., Yokoi, N., Yano, H., Miyazaki, M.,Nakajima, N., Iwanaga, T., Yokoyama, Y., Shibata, T., and Seino, S.(2000) MTABC3, a novel mitochondrial ATP-binding cassette proteininvolved in iron homeostasis. J. Biol. Chem. 275, 17536–17540

57. Krishnamurthy, P. C., Du, G., Fukuda, Y., Sun, D., Sampath, J., Mercer,K. E., Wang, J., Sosa-Pineda, B., Murti, K. G., and Schuetz, J. D. (2006)Identification of a mammalian mitochondrial porphyrin transporter. Na-ture 443, 586–589

58. Paterson, J. K., Shukla, S., Black, C. M., Tachiwada, T., Garfield, S.,Wincovitch, S., Ernst, D. N., Agadir, A., Li, X., Ambudkar, S. V., Szakacs,G., Akiyama, S., and Gottesman, M. M. (2007) Human ABCB6 localizesto both the outer mitochondrial membrane and the plasma membrane.Biochemistry 46, 9443–9452

59. Tsuchida, M., Emi, Y., Kida, Y., and Sakaguchi, M. (2008) Human ABCtransporter isoform B6 (ABCB6) localizes primarily in the Golgi appara-tus. Biochem. Biophys. Res. Commun. 369, 369–375

Subcellular Fractionation and Lysosomal Proteomics

Molecular & Cellular Proteomics 10.4 10.1074/mcp.M110.006403–13

Page 14: This paper is available on line at ...cabm-ms.cabm.rutgers.edu/Classification of subcellular location by... · Classification of Subcellular Location by Comparative Proteomic Analysis

60. Jalil, Y. A., Ritz, V., Jakimenko, A., Schmitz-Salue, C., Siebert, H., Awuah,D., Kotthaus, A., Kietzmann, T., Ziemann, C., and Hirsch-Ernst, K. I.(2008) Vesicular localization of the rat ATP-binding cassette half-trans-porter rAbcb6. Am. J. Physiol. Cell Physiol 294, C579–590

61. Janckila, A. J., and Yam, L. T. (2009) Biology and clinical significance oftartrate-resistant acid phosphatases: new perspectives on an old en-zyme. Calcif. Tissue Int. 85, 465–483

62. Sun, P., Sleat, D. E., Lecocq, M., Hayman, A. R., Jadot, M., and Lobel, P.(2008) Acid phosphatase 5 is responsible for removing the mannose6-phosphate recognition marker from lysosomal proteins. Proc. Natl.Acad. Sci. U.S.A. 105, 16590–16595

63. Einstein, R., and Gabel, C. A. (1991) Cell- and ligand-specific dephosphor-ylation of acid hydrolases: evidence that the mannose 6-phosphatase iscontrolled by compartmentalization. J. Cell Biol. 112, 81–94

64. Della Valle, M. C., Sleat, D. E., Sohar, I., Wen, T., Pintar, J. E., Jadot, M.,and Lobel, P. (2006) Demonstration of lysosomal localization for themammalian ependymin-related protein using classical approachescombined with a novel density shift method. J. Biol. Chem. 281,35436–35445

65. Gasingirwa, M. C., Thirion, J., Costa, C., Flamion, B., Lobel, P., and Jadot,M. (2008) A method to assess the lysosomal residence of proteins incultured cells. Anal. Biochem. 374, 31–40

Subcellular Fractionation and Lysosomal Proteomics

10.1074/mcp.M110.006403–14 Molecular & Cellular Proteomics 10.4