Human Dipeptidyl-peptidase I
GENE CHARACTERIZATION, LOCALIZATION, AND EXPRESSION*

(Received for publication, October 4, 1996, and in revised form, December 20, 1996)

Narayanam V. Rao , Gopna V. Rao and John R. Hoidal Dagger

From the Department of Internal Medicine, Division of Respiratory, Critical Care, and Occupational Medicine, University of Utah Health Sciences Center and Veterans Administration Medical Center, Salt Lake City, Utah 84132

ABSTRACT
INTRODUCTION
EXPERIMENTAL PROCEDURES
RESULTS
DISCUSSION
FOOTNOTES
REFERENCES


ABSTRACT

Dipeptidyl-peptidase I, a lysosomal cysteine proteinase, is important in intracellular degradation of proteins and appears to be a central coordinator for activation of many serine proteinases in immune/inflammatory cells. Little is known about the molecular genetics of the enzyme. In the present investigation the gene for dipeptidyl-peptidase I was cloned and characterized. The gene spans approximately 3.5 kilobases and consists of two exons and one intron. The genomic organization is distinct from the complex structures of the other members of the papain-type cysteine proteinase family. By fluorescence in situ hybridization, the gene was mapped to chromosomal region 11q14.1-q14.3. Analysis of the sequenced 5'-flanking region revealed no classical TATA or CCAAT box in the GC-rich region upstream of cap site. A number of possible regulatory elements that could account for tissue-specific expression were identified. Northern analyses demonstrated that the dipeptidyl-peptidase I message is expressed at high levels in lung, kidney, and placenta, at moderate to low levels in many organs, and at barely detectable levels in the brain, suggesting tissue-specific regulation. Among immune/inflammatory cells, the message is expressed at high levels in polymorphonuclear leukocytes and alveolar macrophages and their precursor cells. Treatment of lymphocytes with interleukin-2 resulted in a significant increase in dipeptidyl-peptidase I mRNA levels, suggesting that this gene is subjected to transcriptional regulation. The results provide initial insights into the molecular basis for the regulation of human dipeptidyl-peptidase I.


INTRODUCTION

Granule-associated serine proteinases are major constituents of polymorphonuclear leukocytes (PMNL),1 cytotoxic lymphocytes, and mast cells, accounting for up to 30% of the cellular protein in these cells. They are involved in many physiologic and pathologic processes. Unlike the more extensively characterized serine proteinases such as trypsin, chymotrypsin, or pancreatic elastase that are stored as inactive zymogens in the secretory vesicles of cells and activated only after secretion into the intestinal lumen, granule-associated serine proteinases of leukocytes and mast cells are stored as fully active enzymes. Nevertheless, based on their known cDNA sequences, these immune/inflammatory cell proteinases are initially translated as zymogens and then processed in several steps including the cleavage of signal peptides and the subsequent removal of short propeptides that typically consist of two amino acid residues (1, 2). The cleavage of the propeptides is unusual in that it occurs at an acidic residue in contrast to most proteinase zymogens that are processed at a basic or, rarely, an aromatic residue. Thus, a major mechanism of control of leukocyte or mast cell granule-associated serine proteinases occurs at the level of dipeptidase activation.

Dipeptidyl-peptidase I (DPP-I, EC 3.4.14.1), a cysteine proteinase, was recently demonstrated to play a requisite role in removing the activation dipeptide from many of the leukocyte and mast cell granule-associated proteinases including human cathepsin G, leukocyte elastase, mast cell chymase and tryptase, and lymphocyte granzymes B and H (3-5). DPP-I, originally called cathepsin C, was discovered when extracts of kidney were found to catalyze the hydrolysis of Gly-Phe-beta -naphthylamide (6). It is a lysosomal enzyme widely expressed in many tissues that is felt to be important in intracellular degradation of proteins. The enzyme was purified from human spleen and characterized as a glycoprotein with a pI of 5.4, a molecular mass of 200 kDa as determined by gel filtration under non-denaturing conditions, and a subunit size of 24 kDa (7).

The strong circumstantial evidence that DPP-I is the central coordinator for activation of many serine proteinases contained in immune/inflammatory cells and that it is differentially expressed in human tissues (8) emphasizes the need for in-depth studies to define factors regulating its expression. A human and a rat DPP-I cDNA have recently been cloned (9, 10), but the reported sequences contained only a portion of the 5'-untranslated regions (UTRs). Moreover, information on gene expression and regulation has not been reported. In the present investigation we describe the structure, localization, and expression of the gene for DPP-I. We also demonstrate its regulation in cytokine-stimulated lymphocytes.


EXPERIMENTAL PROCEDURES

Materials

Maximum strength Nytran membranes were from Schleicher & Schuell, Inc. Multiple tissue Northern blots, ExpressHyb hybridization solution, human spleen total RNA, and Marathon cDNA amplification kits were from CLONTECH. MicroFastTrack kits, cDNA cycle kits, and TA Cloning kits were from Invitrogen. LA PCR kits were from Panvera. TRI REAGENT was from Molecular Research Center Inc. [gamma -32P]ATP (3000 Ci/mmol) and [alpha -32P]dCTP (3000 Ci/mmol) were from Amersham Life Science, Inc. Sequenase DNA sequencing kits were from U. S. Biochemical Corp. RPMI 1640, minimum Eagle's medium, nonessential amino acids, and Cot1 DNA were from Life Technologies, Inc. Defined fetal bovine serum was from Hyclone Laboratories (Logan, UT). Hybrisol VI was from Oncor Inc. Actinomycin D, cycloheximide, and other chemicals not specifically mentioned were high quality grade from Sigma.

Genomic Cloning and Analysis

Poly(A)+ RNA was isolated from human spleen using a MicroFastTrack kit. The first strand cDNA was synthesized with a combination of oligo(dT) and random primers. A portion of the cDNA was amplified using the PCR primers that represent the 5' (743-760 nt) and 3' (complementary to 1429-1446 nt) termini for the rat mature protein coding region (10). The PCR product represented 756-1455 nt of human DPP-I cDNA and was used as a probe to screen a human genomic PAC (1 bacteriophage-derived rtificial hromosome) library (GenomeSystems, Inc.). Phage DNA from the PAC clones was purified by alkali lysis, and the insert was released with NotI followed by digestion with EcoRI. The digested DNA was analyzed by Southern blot hybridization with the human DPP-I cDNA probe used in the library screening, and the relevant DNA fragments were purified from the gel using a Prep-A-Gene kit. To determine the exon and intron organization, fragments of genomic DNA were amplified by PCR and sequenced. To locate intron sequences, the following oligonucleotide primer pairs were selected to amplify overlapping regions spanning the entire length of the human DPP-I cDNA: 1) 13-37 nt and complementary to 855-878 nt, 2) 855-872 nt and complementary to 1438-1455 nt, and 3) 1385-1408 nt and complementary to 1661-1684 nt. The primer pairs used to obtain the 5'-flanking sequence and identify the polyadenylation site were 1) T7 sequencing primer and complementary to 139-162 nt and 2) 1621-1641 nt and SP6 sequencing primer.

To obtain data on the intron size and splice junction site, long and accurate PCR was performed to amplify the fragments of genomic DNA using a GeneAmpTMPCR system 9600. The PCR amplification reaction consisted of an initial denaturation at 94 °C for 1 min, followed by 30 cycles of denaturation at 94 °C for 15 s, annealing at 62 °C for 15 s, and extension at 72 °C for 2 min. Each PCR product was analyzed on an agarose gel, directly subcloned into the pCRTMII vector using the TA cloning kit, and sequenced.

Chromosomal Assignment of Human DPP-I Gene

The genomic plasmid clone was used as a probe for chromosomal localization of DPP-I by fluorescence in situ hybridization. The genomic clone was nick translation-labeled with biotin, hybridized to metaphase chromosomes, and detected with Cy3-conjugated streptavidin. Human metaphase chromosome spreads were prepared by standard procedures and G-banded after trypsin treatment and Wright's staining. Hybridization and detection conditions on metaphase chromosomes were performed as described previously (11). Briefly, the G-banded preparations were destained with a fixative containing methanol and glacial acetic acid (3:1), dehydrated by successive washings in 70, 90, and 100% ice-cold ethanol, and dried at 37 °C. Probe signals were detected with Cy3 conjugate viewed through a triple pass filter using an epifluorescence microscope. The fluorescence in situ hybridization image was overlaid on the G-banded metaphase image to localize the gene.

Determination of Transcription Site

To determine the transcription start site, anchored PCR was performed. Briefly, the first strand cDNA was synthesized at 50 °C using spleen total RNA (1 µg) and a gene-specific primer that was complementary to 855-878 nt. The 5'-end of purified cDNA was (dA)-tailed with terminal deoxynucleotidyltransferase and anchored. The second strand cDNA was synthesized using oligo(dT) containing a 3' rapid amplification of cDNA ends adapter primer. After purification, the double-stranded cDNA was amplified by PCR using a primer pair of an abridged universal amplification primer (Life Technologies, Inc.) and a gene-specific primer complementary to 828-851 nt. The PCR product was re-amplified with an abridged universal amplification primer and a nested gene-specific primer complementary to 751-774 nt. The PCR product was analyzed on an agarose gel, subcloned into PCRTMII vector, and sequenced.

Primer extension was performed to confirm the transcription start site. Briefly, the 18-nt primer complementary to 17-34 nt was end-labeled with [gamma -32P]ATP using T4 polynucleotide kinase. An annealing reaction was carried out with 100 fmol of labeled primer and 25 µg of human spleen total RNA at 58 °C for 20 min. To this reaction, avian myeloblastosis virus reverse transcriptase (1 unit) was added to a 20-µl final volume containing 1 mM dNTPs, 50 mM Tris, pH 8.3, 50 mM KCl, 10 mM MgCl2, 10 mM dithiothreitol, and 0.5 mM spermidine. The extension reaction was carried out at 42 °C for 30 min and terminated by adding 20 µl of loading dye. The primer extension samples were boiled for 10 min and loaded onto a 6% sequencing gel. After electrophoresis, the gel was dried and developed.

Northern Blot Analysis

Multitissue blots were used to determine expression of the DPP-I gene within human tissues. For studies of gene expression in immune/inflammatory cells, U937, PLB 985, or HL-60 cells were grown as described previously (12). Studies of gene regulation were conducted in IL-2-stimulated lymphocytes. Experiments were performed on cells having viabilities of >95% as judged by trypan blue exclusion. Cells were changed to fresh medium before the exposure to possible agonists/modulators at concentrations specified for the indicated periods of time. The cells were harvested at the end of each time period and kept frozen at -80 °C until further analysis. Actinomycin D stock (5 mg/ml) was prepared in 95% (v/v) ethanol. Cycloheximide stock (10 mg/ml) was prepared in phosphate-buffered saline.

Total RNA was prepared by the acid guanidinium thiocyanate phenol/chloroform method (13) and quantified by measuring absorbance at 260 nm. RNA (10 µg/lane) was size-fractionated on 1% agarose, 0.4 M formaldehyde gels containing formamide and transferred to Nytran membranes by capillary action. The RNA was cross-linked to the membrane by exposure to UV light, prehybridized at 68 °C for 30 min, and hybridized with an [alpha -32P]dCTP-labeled probe at 68 °C for 1 h. After several low stringency washes, the blot was washed twice at high stringency and developed.


RESULTS

Organization of Human DPP-I Gene

Overall, the human DPP-I gene spans 3.5 kb and contains two exons separated by 1645 nt of intronic DNA (Fig. 1). The first exon comprises the 5'-UTR followed by 889 nt that encode the signal peptide, propeptide, and partial mature protein region. The second exon contains the remainder of the mature protein-coding sequence of 501 nt, the stop codon, and the 3'-UTR including a polyadenylation signal. The location of the intron was confirmed by sequence analysis. As shown in Fig. 1, the exon-intron boundaries conform to classical splice donor and acceptor consensus sequences (14). The single intron splice site occurred between nucleotides 952 and 953, the first and second nucleotides of the codon for Gly297 in the cDNA sequence, which was indicative of a phase 1 intron. The exon sequence agrees with that determined for the cDNA, indicating that the obtained cDNA using reverse transcription PCR is free of PCR artifacts.


Fig. 1. Schematic representation of the strategy used to determine the structural organization of the human DPP-I gene. The cDNA is depicted indicating the protein coding regions (pre, pro, and mature) and the UTR of the 5'- and 3'-ends. The positions of the 5'-upstream codon (ACC), the translation initiation codon (ATG), the first codon of the mature region (TTG), the stop codon (TAG), and the polyadenylation signal (AATAAA) are shown below the cDNA. The genomic DNA depicts exons as rectangular open boxes and the intron as a solid line, with the nucleotide sequences at the junction of the exon and intron shown below. The invariant dinucleotides of the 5'-donor site and 3'-acceptor site are underlined. The location and orientation of primers used for obtaining the overlapping PCR fragments are illustrated by the closed and open boxes or circles for the sense and antisense primers, respectively. The numbers in parentheses refer to the nucleotide position of the primers in the cDNA sequence.
[View Larger Version of this Image (17K GIF file)]


The cDNA is similar in size and composition to that recently reported for human ileum DPP-I (9) but contains an additional 30 nt of the 5'-UTR, including the transcription initiation site (see below). The DPP-I cDNA sequence spans 1888 nucleotides and includes a 1392-nt open reading frame that encodes for 463 amino acids. This includes a 24-amino acid residue signal peptide, a 206-amino acid residue propeptide region, and a 233-amino acid residue mature enzyme. The coding region of human DPP-I shows 78% identity to the rat at nucleotide and amino acid levels. The mature protein shows 88% identity at nucleotide and amino acid levels with the respective rat sequences. The cDNA sequence differs from that reported for the ileum at nucleotides 276 (C right-arrow G, Leu73 unchanged) and 1440 (G right-arrow A, Pro459 unchanged) in the protein coding region and at nucleotides 1461 (C right-arrow G), 1503 (A right-arrow G), and 1861 (G right-arrow A) in the 3'-UTR. In addition, there is a five-nucleotide (ACTGC) deletion immediately 5' to the poly(A)+ tail in the spleen cDNA when compared with that of the ileum. These five nucleotides (ACTGC) preceding the poly(A)+ tail reported for human ileum DPP-I by Paris et al. (9) are identical to those we observed in the genomic sequence. The basis for this difference is not presently known but may result from the existence of limited genetic variability.

Localization of Human DPP-I Gene by Fluorescence in Situ Hybridization

Of the 20 metaphase cells that were located, all showed Cy3 signals on the long arms of chromosome 11. Fourteen of 20 showed four hybridization signals (one per chromatid, two on each chromosome 11) whereas 6 showed only one signal on one chromosome 11 and two signals on the other chromosome 11. No other chromosomes showed signals with the genomic probe, suggesting a single genomic sequence with high homology to the DPP-I gene locus. Imaging techniques further localized DPP-I to 11q14.1-q14.3 (Fig. 2) with a 92.5% efficiency of hybridization.


Fig. 2. Chromosomal localization of human DPP-I. A photomicrograph of G-banded human metaphase chromosome spreads superimposed with the image of the same metaphase spreads after hybridization with the biotin-labeled plasmid probe containing human DPP-I gene as described under "Experimental Procedures." The arrows show the fluorescent signal on the band q14 of the chromosome 11. To the lower right of the photomicrograph is an ideogram of human chromosome 11 showing the human DPP-I gene location at the region of 11q14.1-q14.3, as indicated by arrows.
[View Larger Version of this Image (26K GIF file)]


Identification of Transcription Initiation Site

Sequence analysis of independent subclones generated by anchored PCR indicated that transcription initiates at the A nucleotide located 63 nt upstream from the ATG that represents the translation initiation codon. Primer extension analysis with the antisense primer positioned 30 nt 5' to the ATG yielded a 34-nt-long product (Fig. 3) calculated to end at the A nucleotide located 63 nt 5' from the ATG and is, therefore, consistent with the results obtained by anchored PCR. The sequence of the region encompassing the transcription initiation site revealed that the A nucleotide is preceded by an invariant C nucleotide and matches the consensus cap signal that has been found in the majority of eukaryotic promoters (15).


Fig. 3. Determination of human DPP-I gene transcription initiation site by primer extension analysis. An 18-nt antisense primer was radiolabeled, annealed to human spleen total RNA, and reverse-transcribed as described under "Experimental Procedures." The products from the primer extension reaction of spleen RNA and positive control RNA were separated on a sequencing gel applied with radiolabeled DNA markers. A single extension product of 34 nt was detected. The transcription initiation site corresponds to an adenosine nucleotide 63 nt 5' from the translation initiation site.
[View Larger Version of this Image (62K GIF file)]


Identification of Putative Regulatory Elements in the 5'-Flanking Sequence of Human DPP-I Gene

The 5'-regulatory region was determined by sequencing the ~3-kb PCR fragment that contained a portion of the first exon and the adjoining upstream region (Fig. 4). Computer analysis of this region revealed several features that are characteristic of promoters. The first 200 nt of the immediate 5'-region relative to the transcription initiation site is GC-rich (65%) as compared with the GC composition (50%) of the whole 5'-region. Further analysis of this region revealed neither a classical TATA box nor a CCAAT box. However, a potential cis-acting DNA element, GC-box/imian virus 40 rotein (Sp1) binding site (position -55 in reverse orientation) (16) was identified. The 5'-region contains recognition sequences for several other transcription factors including three sites for the yclic AMP esponse lement inding rotein (CRE-BP) (17) at positions -519, -950 (reverse orientation), and -953; five sites for CAAT/nhancer inding rotein (C/EBP) (18) at positions -480, -531 (reverse orientation), -660, -732 (reverse orientation), and -665; two sites for NFkappa B/c-Rel (19) at positions -637 (reverse orientation) and -797; and two sites for Oct-1 (20) at positions -897 and -1115 (reverse orientation). Other sites of interest in the 5'-region include binding sites for several cell-specific transcription factors involved in the proliferation and differentiation of hematopoietic cells. These include five yeloid inc inger (MZF1) sites (21) at positions -73 (reverse orientation), -116 (reverse orientation), -349 (reverse orientation), -435, and -1070; nine GATA (22) family binding sites including four GATA-1 sites at positions -76 (reverse orientation), -731 (reverse orientation), -939, and -1055 (reverse orientation), four GATA-2 sites at positions -76 (reverse orientation), -550, -731 (reverse orientation), and -939, and one site for GATA-3 at position -77 (reverse orientation); two sites for aros (IK-2) (23) at positions -691 and -1074 (reverse orientation); one site for the mphoid transcription actor (Lyf-1) (24) at position -1075 (reverse orientation); three sites for v-Myb (25, 26) at positions -563, -654, and -969; one site for the uclear espiratory actor (NRF-2) (27) at position -1054 (reverse orientation); one site for Pbx-1 (28) at position -1111; one site for E1A (arly region of adenovirus)-associated -kDa rotein (p300) (29) at position -357 (reverse orientation); and several sites for CdxA (30). Potential recognition sequences for the Ets (26 ransformation pecific) family of transcription factors c-Ets-1 (31) and ts-ie (Elk-1) (32) are also present in reverse orientation at positions +29, -42, -1052, and -1054 and at position -39, respectively.


Fig. 4. Nucleotide sequence of the 5'-flanking region of human DPP-I gene. The nucleotide sequence contains 1207 nt 5' to the transcription start site (*) and 63 nt 5' to the translation initiation site (), respectively. The putative transcription initiation site of the cDNA sequence is designated +1, with positive and negative numbers proceeding to 3' and to 5', respectively. The first nucleotide of the cDNA sequence of the published study (9) is represented by black-diamond . The sequence was analyzed for regulatory elements that share homology to known transcription factor binding sites using the TFSEARCH program (52). Putative transcription and regulatory elements are underlined and identified above the sequence.
[View Larger Version of this Image (50K GIF file)]


Expression of Human DPP-I mRNA in Tissues

The expression pattern of DPP-I in adult human tissues was determined by Northern blot analysis. The human DPP-I cDNA probe hybridized to a transcript of ~2 kb in all the tissues but with varying intensities. The strongest signal for human DPP-I mRNA was detected in the lung, kidney, and placenta. A signal of moderate intensity was detected in the small intestine, colon, spleen, and pancreas. A low intensity signal was observed in the heart, reproductive organs (testis and ovary), and peripheral blood leukocytes. A weak signal was present in the thymus, prostate, liver, and skeletal muscle. Transcripts were barely detectable in the brain.

Because of our interest in DPP-I as a central coordinator in activating granule-associated serine proteinases, we determined the expression of human DPP-I mRNA in immune/inflammatory cells and their precursors. A representative Northern blot is shown in Fig. 5. Among the fully differentiated cells, the strongest hybridization signal was observed in PMNL and alveolar macrophages. A weak signal was detected in unstimulated lymphocytes and monocytes. Among precursor cells, strong signals were observed in PLB 985, a myelomonoblastic cell line, U937, a myelomonocytic cell line, and HL-60, a promyelocytic cell line.


Fig. 5. Expression of human DPP-I mRNA transcripts in cells of hematopoietic origin. A Northern blot of total cellular RNA (10 µg/lane) from hematopoietic cells hybridized with a 32P-labeled human DPP-I cDNA probe (corresponding to 756-1455 nt) is shown as described under "Experimental Procedures." A hybridization signal at ~2 kb occurred in all the cells but with strikingly different intensities. The bottom panel shows ethidium bromide-stained 28 S and 18 S rRNA on the gel before transfer as assessment of RNA intactness and quantities in each lane.
[View Larger Version of this Image (59K GIF file)]


Regulation of Human DPP-I in Activated Lymphocytes

Because granzymes present in lymphokine-activated lymphocytes are presumably activated by DPP-I that is present in only low levels in unstimulated lymphocytes, we studied the regulation of the human DPP-I gene in lymphocytes stimulated by IL-2. As shown in Fig. 6, low levels of human DPP-I mRNA were present in lymphocytes not exposed to IL-2. When lymphocytes were exposed to IL-2, induction of human DPP-I mRNA occurred as early as 12 h, peaked at 48 h, and then declined by 72 h. Actinomycin D prevented the induction of human DPP-I mRNA in lymphocytes stimulated by IL-2. Treatment of lymphocytes with cycloheximide also prevented the induction of human DPP-I mRNA by IL-2 (data not shown). These results indicate that the induction of human DPP-I mRNA observed in lymphocytes treated with IL-2 most likely occurred at the level of gene transcription and was dependent on protein synthesis.


Fig. 6. Regulation of human DPP-I mRNA transcripts in IL-2-activated lymphocytes. Northern blot of total cellular RNA (10 µg/lane) from lymphocytes treated with IL-2 (1000 units/2 × 106 cells/ml) for the indicated periods is shown. A hybridization signal was barely detectable in unstimulated cells (0 h) and peaked at 48 h in IL-2-stimulated cells.
[View Larger Version of this Image (89K GIF file)]



DISCUSSION

DPP-I is a lysosomal cysteine proteinase differentially expressed in a variety of tissues and thought to play an important role in intracellular protein degradation. To date, it is the only cysteine proteinase that has been demonstrated in PMNL, and recent studies have focused on the importance of this enzyme as a central coordinator for activation of granule-associated serine proteinases contained in PMNL and mast cells. In this investigation, to better understand the molecular basis for the regulation and physiologic effects of DPP-I and to gain insight into its tissue-specific and cytokine-induced expression, we have determined the organization, chromosomal location, and expression of the human DPP-I gene.

DPP-I has been classified as a member of the lysosomal papain-type cysteine proteinase family that also includes cathepsins B, H, L, O, and S. This classification is based on its localization within the cell, acidic pH optimum for enzyme activity, and conserved amino acid sequence with respect to the NH2-terminal and COOH-terminal regions that form the substrate-binding pocket of the enzyme. However, unlike cathepsin B, H, L, O, and S, which are monomeric proteins (molecular mass 20-30 kDa) with endopeptidase activity, DPP-I is an oligomeric protein (200 kDa) with exopeptidase activity. In addition, the overall amino acid sequence homology of DPP-I shows relatively little identity with other members of this group of proteinases.

We now report that the organization of the human DPP-I gene contrasts strikingly with that of the other enzymes contained within the papain group. The human DPP-I gene is of limited size and complexity, existing as a single copy that spans approximately 3.5 kb, contains two exons divided by a single intron, and is expressed as a single transcript. Reports of genes previously described for cathepsin B (33), cathepsin H (34), cathepsin L (35), and cathepsin S (36) emphasize complex structures consisting of multiple exons and introns, some undergoing alternative splicing that gives rise to multiple transcripts that are differentially expressed. Recently, an alignment/phylogeny of the papain superfamily of cysteine proteases was created (37) in which cathepsin B and DPP-I were placed in the same class, appearing to have diverged from the other papain group of sequences before the origin of kinetoplastids. However, the grouping of cathepsin B and DPP-I was not well supported statistically. Based on the results of the current investigation demonstrating the strikingly different genomic organization, we question the grouping of cathepsin B and DPP-I and speculate that rather than having a common ancestral origin with the other mammalian cysteine proteinases of the papain superfamily, DPP-I may have evolved into the class through convergence by selective evolutionary pressure. Also of note, human DPP-I is neither located on the chromosomes of other cysteine proteinase groups (36, 38, 39) in which it is classified nor on chromosomes of granule-associated serine proteinases (40) to which it functions as a processing enzyme.

To begin to address the regulation of human DPP-I gene expression, we analyzed the 5'-flanking sequence for potential upstream regulatory elements. The transcription initiation site is located -63 nt from the translation initiation site and is surrounded by a canonical cap signal sequence. Consensus transcription sequences such as TATA and CCAAT are notably absent in the GC-rich upstream region of the cap site. Eukaryotic promoters lacking a TATA or CCAAT sequence frequently encode proteins or enzymes with housekeeping functions. Most of these constitutively expressed genes exhibit multiple transcription initiation sites distributed over a limited region (41) and multiple Sp1 binding sequences. In contrast, the human DPP-I promoter has a single transcription initiation site and a single Sp1 site. This is similar to the cathepsin S (36), the thrombin receptor (42), and the nerve growth factor receptor (43) genes that are also subject to regulated expression.

The 5'-flanking region of the human DPP-I gene contains putative regulatory elements. Many of these elements (e.g. MZF1, v-Myb, GATA) have been shown to be important in proliferation and differentiation of hematopoietic cells. The promoter region contains T cell-specific transcription factor binding sites (e.g. IK-2/Lyf-1, p300) as well as NF-kappa B recognition sites reported to be involved in the cytokine-stimulated gene expression (44). Further functional analysis of the promoter region will be necessary to determine which of the factors are involved in the regulation of the human DPP-I gene.

It has been reported previously that DPP-I mRNA is widely expressed in all rat tissues (10) and that the relative level of message in different tissues mirrored the protein content (45) as well as the enzyme activity (8). This also appears true for the human, where the level of expression of DPP-I transcripts observed in the current investigation is in close agreement with reports of the distribution of DPP-I enzyme activity in tissues (8) and hematopoietic cells (3, 7). However, there are notable differences in the expression of DPP-I in the human and the rat. In the human, for example, DPP-I is expressed at low or barely detectable levels in the liver and brain, whereas in the rat DPP-I is highly expressed in the liver and moderately expressed in the brain. These results suggest species-specific expression of the enzyme.

Importantly, the pattern of expression of DPP-I transcripts in immune/inflammatory cells is distinct from that observed for granule-associated serine proteinases. Results from the current investigation suggest that DPP-I is expressed at all stages of myeloid cell development. In contrast, mRNA expression of granule-associated serine proteinases is restricted to specific stages of myeloid cell development (46-51). This suggests a role for DPP-I in immune/inflammatory cells that extends beyond that of a processing enzyme for the granule-associated enzymes.

In summary, we isolated and characterized the human DPP-I gene including a 1.2-kb promoter region. The gene contains a single intron and maps to chromosome 11q14.1-q14.3. The putative promoter region has neither consensus TATA nor CCAAT sequences, a characteristic of housekeeping genes, but it appears to be regulated, at least in certain settings. Further studies are needed to determine the basis for this regulated expression.


FOOTNOTES

*   This research was supported by National Institutes of Health Grant HL37615-09 and Veterans Administration Research Services Grant HL07636.The costs of publication of this article were defrayed in part by the payment of page charges. The article must therefore be hereby marked "advertisement" in accordance with 18 U.S.C. Section 1734 solely to indicate this fact.

The nucleotide sequence(s) reported in this paper has been submitted to the GenBankTM/EMBL Data Bank with accession number(s) U79415[GenBank].


Dagger    To whom correspondence should be addressed: Pulmonary Division, Rm. 743A, Wintrobe Bldg., 50 N. Medical Dr., University of Utah Health Sciences Center, Salt Lake City, UT 84132. Tel.: 801-581-7806; Fax: 801-585-3355; E-mail: jhoidal{at}med.utah.edu.
1   The abbreviations used are: PMNL, polymorphonuclear leukocytes; DPP-I, dipeptidyl-peptidase I; UTR, untranslated region; PCR, polymerase chain reaction; nt, nucleotide(s); IL-2, interleukin-2; kb, kilobase(s).

REFERENCES

  1. Salvesen, G., and Enghild, J. J. (1990) Biochemistry 29, 5304-5308 [Medline] [Order article via Infotrieve]
  2. Rao, N. V., Rao, G. V., Marshall, B. C., and Hoidal, J. R. (1996) J. Biol. Chem. 271, 2972-2978 [Abstract/Free Full Text]
  3. McGuire, M. J., Lipsky, P. E., and Thiele, D. L. (1993) J. Biol. Chem. 268, 2458-2467 [Abstract/Free Full Text]
  4. Dikov, M. M., Springman, E. B., Yeola, S., and Serafin, W. E. (1994) J. Biol. Chem. 269, 25897-25904 [Abstract/Free Full Text]
  5. Murakami, M., Karnik, S. S., and Husain, A. (1995) J. Biol. Chem. 270, 2218-2223 [Abstract/Free Full Text]
  6. Gutmann, H. R., and Fruton, J. S. (1948) J. Biol. Chem. 174, 851-858 [Free Full Text]
  7. McGuire, M. J., Lipsky, P. E., and Thiele, D. W. (1992) Arch. Biochem. Biophys. 295, 280-288 [Medline] [Order article via Infotrieve]
  8. Vanha-Perttula, T., and Kalliomäki, J. L. (1973) Clin. Chim. Acta 44, 249-258 [Medline] [Order article via Infotrieve]
  9. Paris, A., Strukelj, B., Pungercar, J., Renko, M., Dolence, I., and Turk, V. (1995) FEBS Lett. 369, 326-330 [CrossRef][Medline] [Order article via Infotrieve]
  10. Ishidoh, K., Muno, D., Sato, N., and Kominami, E. (1991) J. Biol. Chem. 266, 16312-16317 [Abstract/Free Full Text]
  11. Pinkel, D., Straume, T., and Gray, J. W. (1986) Proc. Natl. Acad. Sci. U. S. A. 83, 2934-2938 [Abstract]
  12. Senior, R. M., Campbell, E. J., Landis, J. A., Cox, F. R., Kuhn, C., and Koren, H. S. (1982) J. Clin. Invest. 69, 384-393 [Medline] [Order article via Infotrieve]
  13. Chomczynski, P., and Sachhi, N. (1987) Anal. Biochem. 162, 156-159 [CrossRef][Medline] [Order article via Infotrieve]
  14. Breathnach, R., and Chambon, P. (1981) Annu. Rev. Biochem. 50, 349-383 [CrossRef][Medline] [Order article via Infotrieve]
  15. Bucher, P. (1990) J. Mol. Biol. 212, 563-578 [Medline] [Order article via Infotrieve]
  16. Briggs, M. R., Kadonaga, J. T., Bell, S. P., and Tjian, R. (1986) Science 234, 47-52 [Medline] [Order article via Infotrieve]
  17. Benbrook, D. M., and Jones, N. C. (1994) Nucleic Acids Res. 22, 1463-1469 [Abstract]
  18. Ryden, T. A., and Beemon, K. (1989) Mol. Cell. Biol. 9, 1155-1164 [Medline] [Order article via Infotrieve]
  19. Kunsch, C., Ruben, S. M., and Rosen, C. A. (1992) Mol. Cell. Biol. 12, 4412-4421 [Abstract]
  20. Verrijzer, C. P., Alkema, M. J., van Weperen, W. W., van Leeuwen, H. C., Strating, M. J. J., and van der Vliet, P. C. (1992) EMBO J. 11, 4993-5003 [Abstract]
  21. Morris, J. F., Hromas, R., and Rauscher, F. J. (1994) Mol. Cell. Biol. 14, 1786-1795 [Abstract]
  22. Orkin, S. H. (1992) Blood 80, 575-581 [Medline] [Order article via Infotrieve]
  23. Molnár, A., and Georgopoulos, K. (1994) Mol. Cell. Biol. 14, 8292-8303 [Abstract]
  24. Lo, K., Landau, N. R., and Smale, S. T. (1991) Mol. Cell. Biol. 11, 5229-5243 [Medline] [Order article via Infotrieve]
  25. Biedenkapp, H., Borgmeyer, U., Sippel, A., and Klempnauer, K. (1988) Nature 335, 835-837 [CrossRef][Medline] [Order article via Infotrieve]
  26. Ness, S. A., Marknell, A., and Graf, T. (1989) Cell 59, 1115-1125 [Medline] [Order article via Infotrieve]
  27. Virbasius, J. V., and Scarpulla, R. C. (1994) Proc. Natl. Acad. Sci. U. S. A. 91, 1309-1313 [Abstract]
  28. Lu, Q., Wright, D. D., and Kamps, M. P. (1994) Mol. Cell. Biol. 14, 3938-3948 [Abstract]
  29. Rikitake, Y., and Moran, E. (1992) Mol. Cell. Biol. 12, 2826-2836 [Abstract]
  30. Margalit, Y., Yarus, S., Shapira, E., Gruenbaum, Y., and Fainsod, A. (1993) Nucleic Acids Res. 21, 4915-4922 [Abstract]
  31. Macleod, K., Leprince, D., and Stehelin, D. (1992) Trends Biochem. Sci. 17, 251-256 [CrossRef][Medline] [Order article via Infotrieve]
  32. Hipskind, R., Rao, V., Mueller, C., Reddy, E., and Nordheim, A. (1991) Nature 354, 531-534 [CrossRef][Medline] [Order article via Infotrieve]
  33. Gong, Q., Chan, S. J., Bajkowski, A. S., Steiner, D. F., and Frankfater, A. (1993) DNA Cell Biol. 12, 299-309 [Medline] [Order article via Infotrieve]
  34. Ishidoh, K., Suzuki, K., Katunuama, N., and Kominami, E. (1991) Biomed. Biochim. Acta 50, 541-547 [Medline] [Order article via Infotrieve]
  35. Chauhan, S. S., Popesecu, N. C., Ray, D., Fleischmann, R., Gottesman, M. M., and Troen, B. R. (1993) J. Biol. Chem. 268, 1039-1045 [Abstract/Free Full Text]
  36. Shi, G. P., Webb, A. C., Foster, K. E., Knoll, J. H., Lemere, C. A., Munger, J. S., and Chapman, H. A. (1994) J. Biol. Chem. 269, 11530-11536 [Abstract/Free Full Text]
  37. Berti, P. J., and Storer, A. C. (1995) J. Mol. Biol. 246, 273-283 [CrossRef][Medline] [Order article via Infotrieve]
  38. Wang, X., Chan, S. J., Eddy, R. L., Byers, M. G., Fukushima, Y., Henry, W. M., Haley, L. L., Steiner, D. F., and Shows, T. B. (1987) Cytogenet. Cell Genet. 46, 710-711
  39. Fan, Y. S., Byers, M. G., Eddy, R. L., Joseph, L., Sukhatme, V., Chan, S. J., and Shows, T. B. (1989) Cytogenet. Cell Genet. 51, 996
  40. Jenne, D. E. (1994) Am. J. Respir. Crit. Care Med. 150, (suppl.) 147-154
  41. Reynolds, G. A., Goldstein, J. L., and Bown, M. S. (1985) J. Biol. Chem. 260, 10369-10377 [Abstract/Free Full Text]
  42. Schmidt, V. A., Vitale, E., and Bahou, W. F. (1996) J. Biol. Chem. 271, 9307-9312 [Abstract/Free Full Text]
  43. Sehgal, A., Patil, N., and Chao, M. (1988) Mol. Cell. Biol. 8, 3160-3167 [Medline] [Order article via Infotrieve]
  44. Michael, J. L., and Baltimore, D. (1989) Cell 58, 227-229 [Medline] [Order article via Infotrieve]
  45. Kominami, E., Ishidoh, K., Muno, D., and Sato, N. (1992) Biol. Chem. Hoppe-Seyler 373, 367-373 [Medline] [Order article via Infotrieve]
  46. Hanson, R., Connolly, N., Burnett, D., Campbell, E., Senior, R., and Ley, T. J. (1990) J. Biol. Chem. 265, 1524-1530 [Abstract/Free Full Text]
  47. Bories, D., Raynal, M. C., Solomon, D. H., Darzynkiewicz, Z., and Cayre, Y. E. (1989) Cell 59, 959-968 [Medline] [Order article via Infotrieve]
  48. Zimmer, M., Medcalf, R. L., Fink, T. M., Mattmann, C., Lichter, P., and Jenne, D. E. (1992) Proc. Natl. Acad. Sci. U. S. A. 89, 8215-8219 [Abstract]
  49. Sturrock, A. B., Franklin, K. F., Rao, G., Marshall, B. C., Rebentisch, M. B., Lemons, R. S., and Hoidal, J. R. (1992) J. Biol. Chem. 267, 21193-21199 [Abstract/Free Full Text]
  50. Fouret, P., du Bois, R. M., Bernaudin, J.-F., Takahashi, H., Ferrans, V. J., and Crystal, R. G. (1989) J. Exp. Med. 169, 833-845 [Abstract]
  51. Takahashi, H., Nukiwa, T., Basset, P., and Crystal, R. G. (1988) J. Biol. Chem. 263, 2543-2547 [Abstract/Free Full Text]
  52. Wingender, E. (1994) J. Biotechnol. 35, 273-280 [CrossRef][Medline] [Order article via Infotrieve]

©1997 by The American Society for Biochemistry and Molecular Biology, Inc.