干细胞之家 - 中国干细胞行业门户第一站

 

 

搜索
朗日生物

免疫细胞治疗专区

欢迎关注干细胞微信公众号

  
查看: 353040|回复: 235
go

The story so far: molecular regulation of the heme oxygenase-1 gene in renal inj [复制链接]

Rank: 1

积分
威望
0  
包包
0  
楼主
发表于 2009-4-22 08:16 |只看该作者 |倒序浏览 |打印
作者:Eric M. Sikorski, Thomas Hock, Nathalie Hill-Kapturczak, and Anupam Agarwal作者单位:Department of Medicine and Division of Nephrology, Hypertension, and Transplantation, Department of Biochemistry and Molecular Biology, University of Florida, Gainesville, Florida 32610 4 H# X3 S6 p( @$ O3 A  b8 u( n
                  
- T) M* o9 q3 b/ _, _. Q                  6 J/ U' ^7 Y) \9 W) E
          2 o6 |6 S8 ?- L8 g6 s, T
                        
% B1 ^, u3 S- M            - u4 k# {: c4 F7 D3 ]1 Q5 p% O2 c# X+ }
            
* j) e' v4 G2 {# c' N+ z            
) C% ~9 E) z* j9 L$ x            
1 @8 U1 e1 A8 V9 V2 o  q0 m8 n                      : N0 F* U/ v% f
        . [, T, Q0 y$ X! ^6 k8 ~
        + w' F2 i1 w0 g
        5 Z3 W/ O' _/ i8 O
          【摘要】5 Q! K8 r( P( x7 s! o. U" p. u3 M/ J" ~
      Heme oxygenases (HOs) catalyze the rate-limiting step in heme degradation, resulting in the formation of iron, carbon monoxide, and biliverdin, the latter of which is subsequently converted to bilirubin by biliverdin reductase. Recent attention has focused on the biological effects of product(s) of this enzymatic reaction, which have important antioxidant, anti-inflammatory, and cytoprotective functions. Two major isoforms of the HO enzyme have been described: an inducible isoform, HO-1, and a constitutively expressed isoform, HO-2. A third isoform, HO-3, closely related to HO-2, has also been described. Several stimuli implicated in the pathogenesis of renal injury, such as heme, nitric oxide, growth factors, angiotensin II, cytokines, and nephrotoxins, induce HO-1. Induction of HO-1 occurs as an adaptive and beneficial response to these stimuli, as demonstrated by studies in renal and non-renal disease states. This review will focus on the molecular regulation of the HO-1 gene in renal injury and will highlight the interspecies differences, predominantly between the rodent and human HO-1 genes. ! I+ Q2 }3 D" b$ J% h
          【关键词】 gene transcription oxidant stress heme proteins8 L7 ~8 t4 L9 Q1 N  q
                  HEME OXYGENASE: THE RATE-LIMITING STEP IN HEME DEGRADATION
& m8 N1 c1 m+ Q( A) u
7 N; g1 `& @- |1 Y5 X0 ITHE HEME OXYGENASE ( HO ) ENZYME system catalyzes the rate-limiting step in heme degradation, producing equimolar quantities of biliverdin, iron, and carbon monoxide (CO) ( Fig. 1 ) ( 131, 219, 220 ). Biliverdin is subsequently converted to bilirubin by biliverdin reductase. Two isoforms of heme oxygenase have been characterized: an inducible enzyme, HO-1, and a constitutive isoform, HO-2 ( 131 ). A third isoform, HO-3, that differs from HO-1 but shares 90% amino acid identity with HO-2, has also been described ( 139 ). As products of different genes, HO-1 and HO-2 share roughly 40% amino acid identity ( 131 ) and have different regulation and tissue distribution. HO-1 (32 kDa) is localized to microsomes and HO-2 (36 kDa) to mitochondria. HO-1 is ubiquitously induced in mammalian tissues, whereas HO-2 is constitutively expressed in the brain, testes, endothelium, distal nephron segments, liver, and myenteric plexus of the gut (reviewed in Ref. 4 ). HO-2 may function as a physiological regulator of cellular function, whereas HO-1 plays a cytoprotective role in modulating tissue responses to injury in pathophysiological states ( 231 ). Several recent reviews and editorials have highlighted the biological effects of the reaction product(s), which possess important antioxidant, anti-inflammatory, and antiapoptotic functions as well as the importance of HO-1 as a potent cytoprotective enzyme ( 38, 44, 51, 55, 58, 61, 74, 92, 105, 108, 143, 152, 177, 179, 184, 191, 197, 199, 202, 224, 256 ). The main focus of this article is to provide a review of the current literature on the molecular regulation of HO-1 gene expression, with particular reference to the differential regulation of the human and the mouse HO-1 genes. In this review, heme oxygenase will refer to HO-1 unless otherwise specified.
5 o+ d5 F) h4 ?5 s' P; y& O% [5 P* `$ b& A5 P
Fig. 1. Heme oxygenase (HO)-catalyzed reaction. Heme (iron protoporphyrin IX) is cleaved between rings A and B by heme oxygenase to yield equimolar quantities of iron (Fe 3  ), carbon monoxide (CO), and biliverdin. O 2 and NADPH are required for this reaction. Biliverdin is then converted to bilirubin by biliverdin reductase. M, V, and P: methyl, vinyl, and propionyl groups, respectively. Reproduced from Ref. 4 with permission./ Q) J- q! {6 S9 m
  u5 e4 g$ F2 b% t6 j1 |) k
FUNCTIONAL RELEVANCE OF HO-1 INDUCTION IN RENAL INJURY$ R' n6 r* P8 N# v" N3 Z4 `

/ f5 Y6 e# q" d0 V+ MInduction of HO-1 is an adaptive and beneficial response to acute renal injury secondary to ischemia-reperfusion injury ( 132, 200 ); nephrotoxins (e.g., cisplatin) ( 1, 201 ); glomerulonephritis ( 48, 144, 156, 229 ); renal transplant rejection ( 46, 130 ); and rhabdomyolysis ( 153, 157 ). The first evidence for the protective effects of HO-1 in renal injury in vivo was provided by the studies of Nath et al. ( 153 ) in a rat model of rhabdomyolysis. HO-1 mRNA was induced within 3-6 h after injury in this model, and administration of an HO-1 inhibitor, tin protoporphyrin, worsened renal damage, while prior induction of HO-1 led to a considerable decrease in mortality ( 153 ). HO-1 mRNA is also induced in the kidney as early as 3-6 h after injury in ischemia-reperfusion ( 132, 200 ) and nephrotoxic acute renal failure ( 1, 90, 243 ), models not dependent on filtered heme proteins. Renal tubular induction of HO-1 is protective in these models as well ( 21, 200, 201, 243 ). Modulation of HO-1 expression using chemical inducers, inhibitors, and HO-1 gene delivery also support a functional role for HO-1 expression in ischemia-reperfusion injury in the liver ( 17 ), brain ( 163, 181 ), and heart ( 45, 78 ). The relevance of HO-1 expression is further substantiated by the presence of HO-1 protein in renal tubules in human ischemic acute tubular necrosis ( 172 ).  L' I. e8 N+ S0 Z  ~- b  O
- i) z7 t: ^: j; e
The biological importance of HO-1 was underscored by the development of the HO-1 knockout mouse and the first description of a patient with HO-1 deficiency ( 187, 237 ). Poss and Tonegawa ( 187 ) first generated mice deficient in HO-1 by targeted deletion of a 3.7-kb region including exons 3 and 4 and a portion of exon 5 of the mouse HO-1 gene. Kidneys of (-/-) mice over 20 wk of age showed evidence of iron deposition in renal cortical tubules. A progressive chronic inflammation characterized by hepatosplenomegaly, lymphadenopathy, leukocytosis, hepatic periportal inflammation, and occasionally glomerulonephritis was reported in addition to the iron deposition. These authors also reported that embryonic fibroblasts from the HO-1-deficient animals were more sensitive to oxidant stimuli such as heme, hydrogen peroxide, paraquat, and cadmium ( 186 ). In vivo administration of endotoxin to (-/-) mice (6-9 wk of age) resulted in significantly more liver injury and mortality. Yet et al. ( 242 ) also generated mice deficient in HO-1 and reported the occurrence of severe right ventricular enlargement after chronic hypoxia in the (-/-) mice compared with wild-type mice. The findings of the protective effects of HO-1 in the glycerol model of acute renal injury have been confirmed in HO-1 knockout mice that demonstrate significantly worse renal function and tubular injury with 100% mortality in the HO-1 (-/-) mice compared with HO-1 ( / ) mice ( 157 ). We have also reported that HO-1 overexpression is cytoprotective in cisplatin-induced renal epithelial cell injury and demonstrated that HO-1 (-/-) mice treated with cisplatin develop more severe renal failure with increased apoptosis and necrosis, compared with cisplatin-treated wild-type or heterozygote mice ( 201 ).8 D- Q# |( z2 Q) S5 ?4 b4 U5 C! r

$ Q9 Q$ T. P( R  ]: z7 nThe human patient with HO-1 deficiency ( 237 ) exhibited several phenotypical features similar to those in the HO-1 knockout mouse, including growth failure, anemia, increased iron binding capacity, increased ferritin, tissue iron deposition, lymphadenopathy, leukocytosis, and increased sensitivity to oxidant injury ( 187, 237 ). In HO-1 knockout mice, iron deposition was detected within renal proximal tubular epithelium ( 187 ). Kidney sections of the HO-1-deficient patient also contained multiple foci of iron deposition in the proximal tubular cells ( 172, 237 ). Clearly, comparison of the HO-1-deficient patient with the HO-1 knockout mice has yielded valuable insights as to the role of HO-1 in renal injury. The clinical relevance and the beneficial effects of HO-1 in different settings of renal injury have been discussed in more detail elsewhere ( 86 ).. g0 K2 E# }$ X
) R% O% `" }) u' i$ ~: b
MECHANISMS MEDIATING THE PROTECTIVE EFFECTS OF HO-1 INDUCTION% |# s0 @( X, M* E: U

- [4 s! Y2 Z2 hThe protective effects of HO-1 are mediated through one or more of several potential mechanisms. Increased HO-1 activity results in degradation of the heme moiety, a toxic prooxidant ( 23, 24 ). The reaction also results in the generation of bilirubin, an antioxidant that is capable of scavenging peroxy radicals, inhibits lipid peroxidation, and has recently been shown to protect cells from a 10,000-fold excess of hydrogen peroxide ( 27, 56, 126, 207 ). In essence, the induction of HO-1 results in a shift of cellular redox toward a more antioxidant state rather than a prooxidant milieu. HO-1 induction has been associated with increased iron efflux, and the latter has been suggested as a mechanism for the cytoprotective effects of HO-1 ( 66 ). In addition, ferritin is coinduced with HO-1, allowing safe sequestration of unbound iron liberated from heme degradation ( 22, 25 ). CO, the gaseous product, has vasodilatory effects similar to those of nitric oxide (NO) ( 103, 114, 134 ), as well as antiapoptotic and cytoprotective functions ( 18, 34, 71, 176, 178, 193, 225 ). Recent studies have also demonstrated an important role for the cell cycle regulatory protein p21 in mediating the protective effects of HO-1 expression in cell injury ( 57, 94 ).
9 l- Z. |5 P$ t$ J" p8 p5 Y0 X
! J4 w3 T6 s1 W1 [It should be noted that HO-1 may have a dual role in tissue pathology and is not "therapeutic" in all instances ( 51, 184, 210, 249, 250 ). Each of the products of the reaction can be potentially injurious as well. CO stimulates mitochondrial generation of free radicals and can poison heme proteins ( 252 ). The iron liberated during heme degradation can catalyze free radical reactions, and increased accumulation of bilirubin is associated with kernicterus in neonates ( 52 ). It has been suggested that an appropriate level of HO-1 induction is beneficial, whereas too much HO-1 may, in fact, be a perpetrator of tissue injury ( 184, 210 ). Taken together, the data suggest that optimal levels of HO-1 are critical to determine whether the ultimate effect is one of protection or worsening of tissue injury. By deciphering the underlying molecular mechanism that controls the level of HO-1 enzyme activity, it will be possible to fine-tune HO-1 gene expression in disease states and exploit its use as a therapeutic strategy in the pathophysiology of renal injury.
% H: V. _5 W$ i1 x( m. y8 C$ B$ B! H: ?" ~; @, d4 ]8 j
MOLECULAR REGULATION OF HO-1
6 @% p# z  B$ Q! e7 X
- G, N5 _, t! y6 n! ]The human HO-1 gene is located on chromosome 22q12 ( 116 ) and consists of 5 exons spanning 14 kb. The cDNAs for three mammalian HO-1 genes, including rat ( 148 ), mouse ( 8 ), and human ( 244 ), have been cloned and sequenced, as well as the HO-1 gene in the chicken ( 128 ). The mechanisms underlying HO-1 induction by its multiple inducers are complex, cell and tissue specific, and tightly regulated at the transcriptional level. However, one common denominator for most of the stimuli that upregulate HO-1 is a significant shift in cellular redox ( 19, 227 ). The induction of HO-1 in response to most stimuli tested, including heme, heavy metals, growth factors, NO, oxidized lipids, and cytokines, has been demonstrated to be a consequence of de novo transcription ( 5, 12, 26, 36, 37, 68, 81, 91, 111, 112, 189 ). Consensus binding sites for nuclear factor- B (NF- B), activator protein-1 (AP-1), AP-2, Sp1, upstream stimulatory factor (USF), c-myc/max and interleukin-6 (IL-6) response elements, as well as other transcription factors have been reported in the promoter region of the human HO-1 gene ( 54, 118, 149, 194, 228 ), suggesting a potential role for these factors in modulating HO-1 induction. Both positive and negative regulatory elements have been discovered in the human HO-1 promoter. Positive regulatory regions containing consensus binding sites for AP-1 (-1872), STATx (-1751), c-Rel (-1723), hepatocyte nuclear factor-1 (HNF-1) (-1709), HNF-4 (-1787), and GATA-X (-1803, -1672) have been identified between -1976 and -1655 bp of the human HO-1 promoter ( 214, 215 ). Interestingly, these regions are functional in HepG2 cells but not in HeLa cells in transient transfection studies ( 215 ). A potential cadmium response element (CdRE) (TGCTAGATTT) has been identified at approximately -4 kb 5' relative to the transcriptional initiation site of the human HO-1 gene ( 218 ). Negative regulatory elements (NRE) containing consensus binding sites for NRE boxes [sequences similar to the silencer elements of the chicken lysozyme gene ( 228 )] have been identified between -981 and -412 bp of the human HO-1 promoter ( 215 ). Another negative regulatory region consisting of a polymorphic GT repeat region is also present in the proximal promoter of the human HO-1 gene ( 238 ). A genomic map of the human HO-1 gene with potential regulatory sites is shown in Fig. 2.
4 ^& R" V. G: p7 E, \+ x
) A+ n& X5 }8 e  W" BFig. 2. Genomic structure of the human HO-1 gene. A genomic map of the human HO-1 gene with relevant regulatory regions in the 5'-proximal and -distal promoter as well as important restriction enzyme sites is shown. The location of several transcription factor consensus binding sites identified by previous studies are indicated ( 54, 118, 214, 215, 218, 238 ). StRE, stress-responsive element; CdRE, cadmium-response element; HSE, heat shock element; (GT)n region, variable length ( 15 - 40 ) polymorphic GT repeat sequence; AP-1 and 2, activator protein 1 and 2; USF, upstream stimulatory factor; HIF, hypoxia-inducible factor; SBE, smad-binding element; HNF-1 and -4, hepatocyte nuclear factor-1 and -4; Sp1, specificity protein 1; STATx, signal transducer and activator of transcription; GATAx, GATA binding proteins; NRE, negative regulatory element. Restriction enzymes are abbreviated as follows: N ( Nhe I), R ( Eco R1), X ( Xho I), Xb ( Xba I), P ( Pst1 ), B ( Bam H1). The map is based on the sequence available through GenBank accession no. Z82244 .
" c" P4 [5 P1 d% ?' L/ ]
) ^$ w! d3 Z5 F8 rThe work of Alam and colleagues ( 7 - 10 ) has identified multiple inducer-specific elements, localized within 10 kb of the 5'-flanking region of the mouse HO-1 gene. Specifically, Alam et al. ( 7 - 9, 14, 15 ) have described two distal promoter regions, named E1 and E2 (previously referred to as SX2 and AB1, respectively) at -4.0 and -10 kb, that are required for induction of the mouse HO-1 gene in response to most of the inducers including heme, NO, heavy metals (cadmium), hydrogen peroxide, hyperoxia, LPS, phorbol ester, sodium arsenite, and various electrophiles ( 7 - 9, 14, 15, 75, 76 ). These investigators have proposed that all of these stimuli have a commonality in their activation mechanism, mediated exclusively via E1 and/or E2. Both these regions contain three repeats of a 10-bp sequence [(T/C)GCTGAGTCA] referred to as stress response elements (StRE) with potential similarity to binding sites for the AP-1 family. Overlapping with this sequence is a putative antioxidant and heme-response element, GCnnnGTCA, which resembles binding sites for the v-Maf oncoprotein and the transcription factor NF-E2 ( 93 ). NF-E2-related transcription factor 2 (Nrf2) binds to this sequence ( 14 ).
/ m1 U2 U! a4 Y5 ^) F1 k; o8 X* S* V3 Y; O) i
Based on these studies, we first evaluated a -9.1-kb human HO-1 promoter fragment that contains regions analogous to the StRE sequences in the mouse E1 and E2 sequences ( regions A and B, respectively) ( Fig. 3 A ) ( 87 ). The human HO-1 sequence corresponding to the E1 and E2 regions is also shown in Fig. 3 B. We observed that these regions were only partially responsible for heme- and cadmium-mediated HO-1 gene induction when tested in human renal proximal tubular epithelial and aortic endothelial cells ( 87 ). Unlike the mouse HO-1 gene, our results have demonstrated that regions A and B in the human HO-1 promoter do not respond to other stimuli, such as oxidized lipids, hyperoxia, iron/hyperoxia, hydrogen peroxide, or transforming growth factor- (TGF- ) ( 89 ). In our efforts to mimic steady-state Northern blot induction, we have identified an enhancer region internal to the human HO-1 gene, which, together with the 4.5-kb promoter, recapitulates levels of induction with heme and cadmium ( 87 ). This enhancer region functions in an orientation-independent manner and requires a region between -3.5 and -4.5 kb of the human HO-1 promoter ( 87 ). The enhancer is not responsive to other stimuli, such as hyperoxia, oxidized lipids, hydrogen peroxide, and TGF- ( 87, unpublished observations). In contrast, an analysis of the entire protein coding region of the mouse HO-1 gene in conjunction with portions of the mouse 5'- and 3'-flanking regions by Alam and Den ( 10 ) did not reveal any regulatory elements in response to heme or cadmium. Further studies to characterize this enhancer region that regulates human HO-1 gene expression as well as its function in the context of the larger -9.1-kb human HO-1 promoter fragment, containing the StRE sequences, are in progress in our laboratory. Whether the transcription factor Nrf2, implicated in the regulation of the mouse HO-1 gene, is involved in activation of the human HO-1 enhancer region would be of significant interest.0 I7 L! J6 }3 Z  u! ~
! T+ J* m; b- n$ n
Fig. 3. Analysis of the E1 and E2 regulatory regions of the mouse HO-1 gene with regions A and B of the human HO-1 genes. A : E1 and E2 regions (previously referred to as SX2 and AB1, respectively) of the mouse HO-1 promoter corresponding to regions A and B of the human HO-1 promoter as well as the recently identified internal enhancer ( 87 ) of the human HO-1 gene ( 12.5 kb) are depicted. B : region A of the human HO-1 promoter contains 2 StREs and a CdRE, and region B contains 3 StREs. The mouse E1 and E2 regions each contain 3 StREs., N& f( ~- Y- |3 n" S
1 ]2 G% p# f* S/ `9 g# ]
ROLE OF NRF2 IN THE REGULATION OF THE MOUSE HO-1 GENE
6 Z" X9 v; ]/ p$ o+ m5 z
. Q1 G8 s' @& o# u& KNrf2 belongs to the cap `n' collar (CNC) transcription factor family and is important for the regulation of several oxidant-responsive genes, including mouse HO-1 and -glutamylcysteine synthetase subunit genes ( 14, 161, 235, 248 ). Nrf2 is involved in mouse HO-1 induction by a variety of stimuli, including oxidants, hyperoxia, heme, and cadmium ( 14, 15, 26, 43, 75 ). It has two other members, Nrf1 and Nrf3, both of which are more widely and abundantly expressed ( 39 ). Originally found in erythoid cells ( 142 ), Nrf2 mRNA is ubiquitously expressed in several different murine tissues ( 101, 142 ). Nrf transcription factors typically exist as heterodimers with a smaller family of proteins known as Mafs (for m usculo a po-neurotic f ibrosarcoma) ( 30, 133 ). The regulatory network involving Maf and CNC transcription factors have been recently reviewed ( 145 ). These small proteins share two common structures, a basic leucine zipper (bZIP) and a CNC region, but they lack a transactivation domain. Activation of Nrf2/Mafk DNA binding activity occurs through several different pathways, including transcription, degradation, and activation ( 30, 133, 145, 205 ). Using yeast two-hybrid and coimmunoprecipitation studies, Nrf2 has been shown to dimerize with ATF-4 and bind to the StRE in the mouse HO-1 promoter and activate transcription after stimulation with cadmium ( 85 ).
) ]. C9 t! C. z! h3 g/ k( Z) Z. M( J) j3 ?
Another mechanism for Nrf2 activation is through interaction with CBP [CREB (cAMP-responsive element binding protein) binding protein] ( 107 ). The interaction was shown to take place using two different transactivation domains termed Neh4 and Neh5. While Neh5 was shown to be a common motif found in CNC proteins, Neh4 was discovered to be a novel motif recently identified in the transcription factors p53 and E2F. Nrf2 interacts with Keap1 (Kelch-like erythroid-derived CNC homology-associating protein 1). Overexpression of Nrf2 results in the direct localization of Nrf2 in the nucleus and potent transactivation ( 106 ). Concomitant expression of Keap1 sequesters Nrf2 from the nucleus and represses Nrf2 activity ( 101 ). In summary, Nrf2 is a potent positive regulator of the mouse HO-1 gene and mediates inducer-dependent gene expression.
% B3 x0 R% N: r3 h+ @+ `( [9 H! Y  ]& z2 X
REPRESSION OF HO-1 GENE EXPRESSION' k. y, o0 x$ ?. {9 E

( b  ~' x* K7 y. W1 DRole of Bach1
+ w3 W! y1 T7 i9 P& \4 S1 y8 |( s; u: A9 I
4 Z3 T  E! f6 i" Y$ cOne antagonist to Nrf2 activation of HO-1 is the transcription factor Bach1 ( 31, 171, 180 ). Bach1 is a bZIP protein that interacts with the small oncoproteins of the Maf family ( 180 ). Bach1 maps to human chromosome 21q22.1 and is expressed as a 5.8-kb transcript in several tissues including the kidney ( 31 ). Bach2 is restricted to B lymphocytes and the brain ( 31 ). Both Bach1 and Bach2 act as transcriptional repressors, but Bach1 can function as a transcriptional activator in erythroid cells ( 180 ). Bach1 heterodimerizes with MafK and binds to multiple Maf recognition sites (MARES) in the mouse HO-1 promoter, repressing gene expression ( 209 ). Recent studies have demonstrated that Bach1 also participates in hypoxia-inducible repression of the human HO-1 gene ( 113 ). Sun et al. ( 209 ) have shown that HO-1 is constitutively expressed at higher levels in many tissues of Bach1-deficient mice, suggesting that Bach1 acts as a negative regulator of transcription of the mouse HO-1 gene. It has been proposed that Bach1 prevents accessibility of enhancers to Nrf2 by binding to them. On heme stimulation, Bach1 is displaced, allowing binding of Nrf2 to its DNA binding sites and consequent transcriptional activation to increase HO-1 expression for heme degradation ( 209 ). The roles played by Bach1 and Nrf2 in mouse HO-1 gene regulation are very similar to the roles of the lac repressor and activator with the lac operon ( 209 ). Bach1 DNA binding activity is negatively affected by the binding of heme to cysteine-proline motifs located in the COOH-terminal region, thereby releasing repression on transcription by Bach1 ( 169 ).
# }7 e  W2 w+ d& r2 ^5 ]8 W0 O9 p. n5 p5 ^: I" J4 Z, B
Length Polymorphisms in the GT Repeat Region' Q1 r+ i. u, r

+ o: V7 D! C0 ^- K1 S& d! yA (GT)n repeat region that functions as a negative regulatory region is located between -198 and -258 of the human HO-1 promoter and is absent in the mouse HO-1 gene. Length polymorphisms of this region vary between subjects and correlate with disease activity in patients with emphysema ( 238 ), coronary artery disease ( 42, 104 ), and vascular restenosis after balloon angioplasty ( 63 ). Individuals with shorter repeats (
" c# L  T2 G4 a. j  P
; O+ ?" I6 {) d+ a6 {It is interesting to note that a protective gene such as HO-1 has developed inherent genetic mechanisms to remain repressed in normal conditions. Kitamuro et al. ( 113 ) have suggested several possible explanations for the physiological implications of HO-1 repression. HO-1 repression reduces energy expenditure because the HO-1 reaction consumes oxygen and NADPH during heme degradation. In addition, as discussed before, the excessive generation of byproducts of the HO-1 reaction, namely, CO, iron, biliverdin, and bilirubin, can have potential harmful effects. It is therefore possible that the Bach1 and the (GT)n repeat region function as repressors of HO-1 expression to prevent undesired excess release of these products in normal physiological states.
. I# J0 y2 L  I0 ]8 s" K
. z: a) S( v, I5 N, A2 eDIFFERENCES BETWEEN HUMAN AND MOUSE HO-1
3 X2 o4 ]  @' @% q0 a; X9 w6 X" x$ q1 ?5 [5 @. l
Studies from our laboratory as well as others demonstrate significant differences in the regulation of the human vs. the mouse HO-1 gene and are summarized in Table 1. For instance, the (GT)n repeat identified in the human HO-1 promoter is not present in the mouse HO-1 gene ( 238 ). A potential CdRE has been identified in region A of the 5'-flanking region of the human HO-1 gene; however, the CdRE in the mouse HO-1 gene is immediately downstream of this region and is a StRE, which is a binding site for Nrf2 and its partners ( 15, 218 ). Our recent studies have identified an internal enhancer in the human HO-1 gene that regulates induction by heme and cadmium ( 87 ), while the E1 and E2 regions of the mouse HO-1 promoter exclusively mediate gene expression ( 14, 15 ). The E1 region in the mouse HO-1 promoter regulates responsiveness to hyperoxia, hydrogen peroxide, and oxidized lipids ( 44, 119, 122 ). However, the analogous region in the human HO-1 promoter is not responsive to these stimuli ( 68, 89 ). We have identified a regulatory region between -9.1 and -11.6 kb of the human HO-1 promoter by DNase I hypersensitivity studies and promoter deletion analysis that, at least in part, mediates oxidized lipid-inducible HO-1 gene expression ( 89 ). Hypoxia is a potent inducer of HO-1 in rat, bovine, mouse, and monkey cells but is a repressor in human cells ( 113, 120 ). In addition, HO-1 has also been referred to as heat shock protein 32 based on its inducibility by heat shock in rodent cells ( 198 ). However, heat shock does not induce HO-1 in human cells ( 174, 194, 197 ). Cytokines such as IFN- induce HO-1 in rodent cells but not in human cells. On the contrary, IFN- represses human HO-1 gene expression ( 197 ). The interspecies differences in the regulation of the human and mouse HO-1 genes will be important considerations for the ultimate use and development of molecular therapies from the "bench to the bedside."
- M  Q( |: J' F  m, ]5 y' ~, m  D2 O  B, C1 u
Table 1. Differences between the regulation of mouse and human HO-1 genes
" R0 o* Y; S# C! w& l" r5 d* u/ L6 U( z6 M9 ~* S
FACTORS RESPONSIBLE FOR HO-1 INDUCTION IN RENAL INJURY
0 }* d7 Y+ m4 Z
6 F  I. u$ f6 }5 Q2 n: h+ G. tThe protective effects of HO-1 activity were first recognized due to its dramatic induction after exposure to a wide variety of injurious stimuli ( 19, 111, 206 ). Inducers of HO-1 include heme; ultraviolet A radiation; hydrogen peroxide; cytokines (IL-1, IL-6, IL-10, TNF-, IFN- ); endotoxin; growth factors (PDGF, TGF- ); heavy metals; oxidized lipids; shear stress; hyperoxia; NO; NO donors; angiotensin II; glucose deprivation; and others ( 2, 5, 6, 12, 13, 15, 16, 27, 35, 37, 40, 41, 49, 59, 60, 68, 69, 79, 81, 83, 88, 95, 100, 115, 119, 121, 135, 156, 221, 232 ). Several of these stimuli, such as heme, nephrotoxins (e.g., cadmium), NO, growth factors, cytokines, and angiotensin II play an important role in the pathophysiology of acute renal failure, as shown in Fig. 4. It has been proposed that these factors activate the HO-1 gene, leading to increased HO-1 enzyme activity and a protective response in renal injury. Figure 5 is a representative Northern blot showing the effects of these stimuli on the induction of HO-1 mRNA in human renal proximal tubule cells. In addition to heme and cadmium, hydrogen peroxide, NO, TGF-, curcumin (the active ingredient in the food spice turmeric), and oxidized lipids also increase HO-1 mRNA expression ( Fig. 5 ). Of these stimuli, heme, cadmium, and NO are the most potent HO-1 inducers, whereas cytokines in combination ( Fig. 5 ) or individually (data not shown) have no effect in these cells. The following discussion will focus on the regulation of HO-1 gene expression by some of these stimuli due to their importance in renal injury and will also include comparisons between human and mouse HO-1 gene regulation.  c3 m( ^6 P/ X" J$ @

* B/ Z( _$ i; F3 J9 Q- uFig. 4. Inducers of HO-1 relevant to the pathogenesis of renal injury. Several stimuli including heme, hypoxia, nitric oxide (NO), cytokines, endotoxin, growth factors, and angiotensin II (ANG II) are implicated in the pathogenesis of renal injury that occurs in the setting of ischemia-reperfusion, nephrotoxins, sepsis, rhabdomyolysis, transplant rejection, and glomerulonephritis. These stimuli are potent HO-1 inducers, and it is proposed that the induction of HO-1 by these factors provides an adaptive and beneficial response in renal injury.
. `% j. }) f* Q- b$ m8 g8 m6 v% d4 c
Fig. 5. Induction of HO-1 mRNA by multiple stimuli in human renal proximal tubule cells. Confluent human renal proximal tubular cells were incubated with control (PBS), hemin (5 µM), cadmium (10 µM), hydrogen peroxide (H 2 O 2; 200 µM), mercuric chloride (mercury; 5 µM), spermine NONOate (NO; 0.5 mM), TGF- (2 ng/ml), curcumin (8 µM), TNF- (10 ng/ml) IL-1 (2 ng/ml) IFN- (5 ng/ml), and oxidized lipids [13-hydroperoxyoctadecadienoic acid (13-HPODE), 20 µM]. Cells were exposed to the stimuli for 4 h, except for mercury, spermine NONOate, H 2 O 2, and 13-HPODE treatments, whereby cells were treated for 1 h followed by the addition of media without stimuli for an additional 3 h.
% |) C6 S) u+ R
& e# W+ [  q4 z- m% YHeme3 j% O! i+ C! g4 Y6 A  i

9 d# T" j6 Y& R. p4 G+ mHeme serves as the prosthetic moiety for heme proteins, such as hemoglobin, myoglobin, cytochromes, prostaglandin endoperoxide synthase, NO synthase (NOS), catalase, peroxidases, respiratory burst oxidase, and pyrrolases, which play an important role in the maintenance of critical functions that include oxygen delivery, mitochondrial respiration, and signal transduction ( 86, 155, 184 ). Interestingly, heme serves not only as a substrate for HO-1 but also stimulates HO-1 gene transcription both in vivo and in cultured cells ( 12, 13, 153, 183 ). In models of renal injury, increased levels of heme are observed in the kidney before the induction of HO-1 ( 1, 132, 200 ). The source of this heme is attributed to either excessive filtration of heme proteins as would occur in the setting of rhabdomyolysis ( 153, 157 ) or from the destabilization of intracellular heme proteins (e.g., cytochromes) in settings of ischemia-reperfusion and nephrotoxin-induced renal injury ( 1, 24, 200 ). Heme is lipophilic and damages multiple cellular targets, including lipid bilayers, mitochondria, cytoskeleton, nuclei, and several intracellular enzymes ( 155 ). While high concentrations of heme are injurious, low concentrations of heme lead to the induction of HO-1 and serve as a cytoprotective response.! C9 L, }: N+ l( _  |  Z

) j  g; R7 q9 J9 k: S. J; T) AHeme binds specifically to Bach1 and regulates its DNA-binding activity ( 169 ). Ogawa et al. ( 169 ) have proposed the following model for transcriptional regulation by heme through Bach1. In the presence of low heme concentrations, genes are repressed through the interaction of Bach1 with MARE sequences, whereas higher concentrations of heme inactivate the binding of Bach1, allowing access of transcription factors such as Nrf2 to interact with the MARE sequences ( 169 ), which in turn allows gene activation. Whether a similar mechanism operates in the transcriptional regulation of the human HO-1 gene by heme remains to be elucidated.
0 w0 y6 I6 I4 V2 O& c1 p( ?' X* a" |* u9 V
In the mouse HO-1 gene, the E1 and E2 regions are required for heme-mediated HO-1 induction ( 7 - 9 ). Recent studies have identified Nrf2, MafG, ATF3, as well as Jun and Fos family members as StRE binding proteins in nuclear extracts from immortalized rat proximal tubular cells exposed to heme ( 11 ). In these studies, heme did not directly increase Nrf2 transcription but rather decreased the rate of Nrf2 degradation ( 11 ). It was proposed that the expression of Nrf2 in unstimulated cells was curtailed by rapid degradation and HO-1 gene expression is maintained at a low level by the binding of repressor proteins (e.g., Bach1) to the StRE sites in the E1 and/or E2 regions ( 11, 209 ). Heme stimulation leads to decreased Nrf2 degradation, allowing for accumulation of Nrf2 in the nucleus, where it heterodimerizes with partners such as MafG. These heterodimers can lead to the displacement of the repressor proteins bound to the HO-1 regulatory regions and increase HO-1 gene transcription ( 11 ).4 D- \0 d- b8 m% U, ~  I5 k
5 k  Y& _* C5 J3 s" c5 B: i( O
Our laboratory has evaluated the regulation of the human HO-1 gene by heme in human renal proximal tubular and aortic endothelial cells ( 87 ). Heme stimulation (5 µM) causes a 20- to 30-fold induction of HO-1 mRNA at 4-h incubation. Such induction occurs via direct increases in de novo gene transcription and is not dependent on increased mRNA stability, findings consistent with previous studies in other cell types ( 5, 12 ). Based on the identification of heme-responsive elements in the mouse HO-1 gene, we first evaluated multiple human HO-1 promoter constructs up to -9.1 kb from the transcriptional start site. These constructs contained the StRE sequences similar to those described in the E1 and E2 regions of the mouse HO-1 gene ( Fig. 3 ). We observed only a partial response of the reporter gene after heme stimulation. The levels of the reporter gene did not correlate with steady-state Northern blot analysis levels of HO-1 induction with heme, suggesting that additional regulatory sequences are required. In an effort to mimic endogenous stimulus-dependent levels of HO-1 induction, we evaluated the entire 12.5-kb of the human HO-1 gene, including introns and exons, in conjunction with a -4.5-kb human HO-1 promoter and observed significant heme- and cadmium-mediated induction, suggesting the presence of an internal enhancer ( 87 ). Similar results were observed with two different reporter genes, human growth hormone and luciferase. The internal enhancer functioned as a true enhancer because it functioned in both orientations (5'-3' and 3'-5') ( 87 ). The enhancer did not function in the context of a heterologous thymidine kinase promoter and required a region between -3.5 and -4.5 kb of the human HO-1 promoter for transcriptional activation ( 87 ). Further studies to delineate the important regulatory sequences within this internal enhancer are in progress.# u! K2 U6 t7 U0 J
5 o8 v0 b; A& I1 n. y2 H
Cadmium* G2 F. _0 }* j9 H" R, d. R
, b. A( P' e0 V1 C
Occupational exposure, such as working with cadmium-containing pigments, plastics, glass, metal alloys, and electrode material in nickel-cadmium batteries, and nonoccupational exposure, such as food, water, and cigarette smoke, contribute to the buildup of cadmium in the body. Cadmium, absorbed either through the lungs or the gastrointestinal tract, accumulates mainly in the kidneys and liver and has a long biological half-life of 15 yr. Chronic long-term exposure leads to renal dysfunction with slow progression to stage 5 chronic kidney disease. In the kidney, the proximal tubular cells are the predominant site for cadmium accumulation and cell injury. After absorption, cadmium is bound to apoprotein, metallothionein, and the cadmium-protein complex is filtered through the glomerulus into the urinary space where it becomes endocytosed by the proximal tubule cells and degraded by the lysosomes, resulting in the release of the cadmium and consequent cellular damage. Cadmium-induced cell injury is mediated via the generation of reactive oxygen species, lipid peroxidation, protein crosslinking, DNA damage, and alteration of intracellular calcium.. I/ O8 p. v3 d& A2 N0 b

- P- w) {6 P) I2 u" R, Y, `Cadmium is a potent inducer of HO-1 gene expression ( 12, 15, 217, 218 ). Such induction may represent a cellular defense against cadmium-mediated injury ( 218 ). HO-1 induction by cadmium occurs via transcriptional activation ( 12 ). Takeda et al. ( 218 ) have identified a 10-bp sequence (CdRE), TGCTAGATTT, at approximately -4.0 kb of the human HO-1 promoter that confers cadmium-mediated induction in HeLa cells. The GC dinucleotides and the G residue in the CdRE are essential for cadmium-mediated gene activation ( 218 ). Immediately downstream of this CdRE sequence is a StRE sequence that is also present in the E1 region of the mouse HO-1 gene ( Fig. 3 B ). Mutational analysis of the StRE sequence in the human HO-1 promoter did not affect cadmium-mediated induction ( 218 ). The CdRE of the human HO-1 gene was not responsive to other HO-1 inducers, including heme, sodium arsenite, and cobalt protoporphyrin, and metals such as zinc ( 218 ). The CdRE is distinct from the metal-responsive element identified in the human metallothionein gene, which is responsive to both cadmium and zinc ( 217 ). The CdRE is also involved in the induction of human HO-1 by tobacco smoke and hydrogen peroxide in human monocytic cells ( 64 ). The proteins binding to the human CdRE have not as yet been identified. It should be noted that only a threefold increase in reporter activity was observed in HeLa cells, levels consistent with our studies in human renal proximal tubular cells with the -4.5-kb construct ( 5 ). Similar to heme, we have reported that cadmium-mediated HO-1 induction also requires the internal enhancer to recapitulate steady-state Northern blot analysis level induction of the endogenous HO-1 gene in human cultured cells ( 87 ).8 f  _' `4 ~( q! S5 u6 o
+ X/ Q! s+ \. {
In the mouse HO-1 gene, the E1 region is necessary for HO-1 gene activation by cadmium and involves the p38 MAPK pathway and Nrf2 ( 15 ). However, significant activation of a 15-kb mouse HO-1 promoter containing the E1 and E2 regions is seen only in MCF7 cells, a human mammary epithelial cell line ( 15 ). A mechanism involving Nrf2 has been proposed for cadmium-mediated activation of the mouse HO-1 gene in mouse hepatoma cells ( 205 ). Cadmium increased the half-life of Nrf2, and this increase in Nrf2 was via the ubiquitin-proteasome pathway, because proteasomal inhibitors enhanced Nrf2 expression ( 205 ). Studies in the rat HO-1 gene have demonstrated a role for USF in cadmium-mediated HO-1 induction in rat glioma cells ( 129 ).+ @. Q  j+ D, ?) ?: s; X' e

: V5 k2 a0 B: B, H% h/ ?/ SRecent studies by Zhang et al. ( 253 ) have generated a transgenic mouse using a 15-kb mouse HO-1 promoter linked to the luciferase gene to allow for in vivo spatiotemporal transcription patterns of HO-1 gene expression in living animals. Injection of cadmium into these mice led to a significant, dose-dependent increase in luciferase activity in the liver and kidney ( 253 ). A modest increase in luciferase gene expression was observed after heme was administered intraperitoneally, but not when heme was given intravenously to these mice, possibly due to the formation of heme-serum protein complexes ( 253 ). The ability to monitor changes in gene expression in vivo in living animals offers a novel approach to test and identify potential therapeutic agents previously determined by in vitro molecular studies.
% u% w3 E9 ]6 \! s
' V- y) _' C% U  f$ k" K5 e! L4 NOther nephrotoxins such as cisplatin and mercury also induce HO-1 in renal tubular cells ( 1, 154 ). In cisplatin-induced renal injury, the induction of HO-1 is preceeded by increases in renal heme content, suggesting that heme may be the stimulus responsible for mediating HO-1 induction ( 1 ). Recent studies have demonstrated that cytokines such as TNF- and IL-1 and growth factors such as TGF- are increased in a mouse model of cisplatin-induced renal injury ( 188 ). It is possible that these mediators also contribute to HO-1 induction in this model. The molecular mechanism(s) and the regulatory sequences controlling HO-1 induction by cisplatin have not been delineated.
7 `/ B) |4 }$ x" I' z. M+ i) A7 `" u; b* e% m  r
The role of NO in ischemia- and nephrotoxin-induced renal injury has been recently reviewed ( 158, 168, 245 ). NO is generated from L -arginine by a group of heme-containing enzymes, NOS. NO plays a dual role in acute renal injury because it can both attenuate or exacerbate renal injury, depending on a balance between beneficial hemodynamic effects and cytotoxicity as well as the site and rate of NO production and the chemical fate of the NO produced ( 167, 168, 245, 246 ). The inducible isoform of NOS (iNOS) generates significant amounts of NO in response to cytokines and endotoxin and is critical to NO-dependent toxicity in both in vitro and in vivo models of renal ischemia-reperfusion injury ( 167, 168 ). Inhibition of iNOS, using iNOS antisense oligonucleotides, attenuates renal ischemia-reperfusion injury ( 167 ), which is further corroborated by studies in iNOS knockout mice ( 124, 125 ). NO induces cell injury through multiple mechanisms including generation of peroxynitrite, nitrosylation of thiols, and impairment of iron-sulfur clusters of proteins ( 28 ). Numerous studies have reported the capacity of NO and NO donors to dramatically induce HO-1 in diverse cell types including endothelial, smooth muscle, renal tubular, and mesangial cells ( 16, 33, 41, 49, 59, 69, 79, 81, 123, 135, 212 ). Studies have also demonstrated that NO-mediated HO-1 induction attenuates cytokine- and oxidant stress-induced cytotoxicity ( 185 ). Most studies have implied that induction of HO-1 by distinct NO donors occurs predominantly via transcriptional mechanisms ( 59, 81 ) and, in part, through increased HO-1 mRNA stability ( 33, 81 ), and represents an adaptive response to the harmful effects of NO.
% d# F! U# l& I2 }3 L9 U/ |: a" @
Nuclear run-on studies have confirmed that de novo transcription is responsible for the NO-mediated activation of the HO-1 gene in rat vascular smooth muscle cells ( 59 ). However, the regulatory regions that are responsible for NO-mediated HO-1 gene induction have not been identified. Marquis and Demple ( 135 ) evaluated a 4.7-kb HO-1 promoter construct in IMR-90 cells (a human embryonic lung fibroblast cell line) and HeLa cells and observed a 0 q( L1 ?7 S0 ?% n) o) H- p7 s
- W2 Z( r3 g7 G" u* [
Studies have also demonstrated that HO-1 mRNA is dramatically stabilized after NO exposure in IMR-90 cells ( 33 ). Iron chelation with deferoxamine and the antioxidant N -acetylcysteine (NAC) block NO donor-mediated HO-1 induction in LLC-PK 1 (porcine renal proximal tubular cells) and in rat aortic smooth muscle cells ( 81, 123 ). The NO scavenger carboxy-2-phenyl-4,4,5,5-tetramethylimidazoline-1-oxyl-3-oxide completely blocks NO-mediated HO-1 induction ( 123 ), whereas cGMP is not involved in NO-mediated HO-1 induction ( 123, 147 ). The reactive nitrogen compound peroxynitrite, formed by the interaction of NO with superoxide anion, induces HO-1 in bovine vascular endothelial cells but not in LLC-PK 1 cells ( 70, 123 ). Incubation of rat smooth muscle cells with a mixture of cytokines (IL-1 and TNF- ) increases nitrite production as well as induced HO-1 expression ( 59 ). Inhibition of iNOS using N G -methyl- L -arginine attenuated both the production of nitrite as well as the induction of HO-1, suggesting that endogenously released NO can also stimulate HO-1 gene expression ( 59 ).$ I9 p- s% c2 F# l0 v' @: S

$ _* C9 R. C! X! vMost previous studies have used NO donors such as sodium nitroprusside (SNP), S -nitroso- N -acetylpenicillamine (SNAP), and 3-morpholinosydnonimine (SIN-1) to evaluate effects of NO on gene expression. However, these donors have significant drawbacks. For example, SNP releases cyanide and iron and SIN-1 releases NO and superoxide simultaneously. Levels of nitrite production with these donors also do not correlate with the levels of HO-1 induction. For instance, SNP releases the lowest amount of nitrite ( 10 µM) compared with SIN-1 ( 40 µM), but SNP is the most potent inducer of HO-1 mRNA ( 79 ). Differential effects of NO donors have been observed in transient transfection studies as well. Hara et al. ( 79 ) tested the 4.5-kb human HO-1 promoter and observed that treatment with SNP, but not with S -nitroso- L -glutathione (GSNO) or SIN-1, increased the expression of the reporter gene through the CdRE of the human HO-1 gene, suggesting that SNP induces HO-1 mRNA expression through a mechanism different from that for GSNO or SIN-1. To more directly study the effects of NO on HO-1 gene expression, NO gas or NO-releasing compounds belonging to the diazeniumdiolate family (called NONOates) have been used as sources of bioactive NO ( 67, 135 ). The commonly used NONOates include diethyltriamine/NO and spermine NONOate. Diethyltriamine/NO is a slow NO generator, whereas spermine NONOate mimics a burst release of NO. The by-products of these NO generators do not affect HO-1 mRNA expression ( 33 ). Consistent with the observations in other cell types, we have observed significant HO-1 mRNA induction by spermine NONOate in human renal proximal tubular cells ( Fig. 5 ).
0 z( r0 e6 G; [" R9 k4 H* @: A6 }4 S- u% m; Z
Growth Factors) E: @$ q! W! R  T' {+ {7 R2 d
  c; U3 Q& b/ p$ W4 c; O2 Y
TGF-, a member of the TGF- superfamily, is a regulatory cytokine implicated in a variety of kidney diseases where, on the one hand, it promotes scarring and proinflammatory events and, on the other hand, stabilizes and attenuates tissue injury ( 204 ) through the activation of cytoprotective proteins, including HO-1 ( 109, 136 ). Increased TGF- is the final common pathway in the pathogenesis of kidney disease by several factors including angiotensin II, hypoxia, high glucose, auto-antibodies, immune complexes, advanced glycosylation end products, and PDGF ( 32, 166 ). In response to injury, TGF- and other growth factors are released via autocrine and/or paracrine mechanisms to maintain cellular homeostasis. While chronic elevation of TGF- plays an important role in the progression of renal diseases ( 195 ), TGF- also has beneficial effects. TGF- has been shown to increase the expression of HO-1 in human retinal pigment epithelial cells ( 117 ), human renal proximal tubular epithelial cells ( 88 ), and human pulmonary epithelial cells derived from a lung cell carcinoma (A549 cells) ( 164 ), as well as in the bovine choroid fibroblasts ( 117 ). However, TGF- does not induce HO-1 in all cell types, including HeLa, HEL, or bovine corneal fibroblasts ( 117 ). Interestingly, in an LPS-induced rat model of endotoxemia, as well as in IL-1 -treated cultured rat vascular smooth muscle cells (i.e., HO-1 is preinduced), TGF- conferred a negative effect on HO-1 and contributed to the beneficial effects of TGF- in a model of endotoxic shock ( 182 ). It is speculated that TGF- -mediated HO-1 induction may counteract the negative effects of TGF- by affecting cell proliferation, apoptosis, and deposition of extracellular matrix.
. k4 U2 t  P' b/ L- N$ _: t; S. |6 U  l3 t4 e
The molecular regulation and signaling pathway(s) involved in TGF- -mediated HO-1 induction are incompletely understood. TGF- initiates signaling through interactions with TGF- type I and type II receptors, which can subsequently activate a group of structurally related proteins called Smads ( 136 ). Thus far, Smad proteins are the only known TGF- receptor substrates and signal transducers involved in many signaling responses induced by TGF- ( 136 ). The molecular mechanism of TGF- signal transduction through Smads has been studied extensively in cancer models. Smad6 and Smad7, the anti-Smads, inhibit activation of TGF- -responsive genes such as human plasminogen activator inhibitor-I (PAI-1) and collagen ( 20, 62, 84, 151, 254 ). Our previous studies have demonstrated that TGF- induced not only HO-1 but Smad7 as well in human renal epithelial cells ( 88 ). Furthermore, overexpression of Smad7, but not Smad6, inhibited the induction of the endogenous HO-1 gene ( 88 ).% {6 T3 l$ A# v7 F

) T7 M* ^/ w- {) W8 l2 ]( `* VThe ability of several genes to respond to members of the TGF- family requires the presence of one or more Smad binding elements (SBE) ( 82, 150, 230 ). Putative SBEs have been described in the human PAI-1 promoter, which is responsive to TGF- ( 53, 208 ). A palindromic sequence, GTCTAGAC, has been described as the Smad3-Smad4 binding element. However, optimal Smad binding is reportedly achieved with a 5-bp sequence CAGAC ( 196, 251 ). The original palindromic sequence, which may have resulted from dimerization of recombinant Smads used in oligonucleotide selection experiments ( 136, 251 ), is not present in the human HO-1 gene. However, computer analysis reveals a consensus sequence, GTCTATAC, located at -5.7 kb in the human HO-1 promoter. To identify TGF- -responsive cis -acting regulatory elements, we have evaluated several human HO-1 promoter fragments in transient transfection studies in human renal proximal tubular cells. An 11.6-kb HO-1 promoter construct elicited an approximately twofold increase in reporter activity, which was attenuated by cotransfection with Smad7 (Hill-Kapturczak N and Agarwal A, unpublished observations).
0 w0 y8 Y. M7 l" Z- z# N6 @" o% q% }, v# @* I
It has also been suggested, however, that TGF- can signal through MAPK pathways, which may be independent of Smad proteins ( 47, 136, 247 ). It was demonstrated, using a chemical inhibitor of p38 MAPK (SB-203580) as well as transfection of a dominant-negative p38 MAPK mutant, that p38 MAPK may be responsible for transducing TGF- signaling and stimulating HO-1 gene expression in A549 cells ( 164 ). However, SB-203580 did not prevent induction of HO-1 by TGF- 1 in human renal proximal tubular cells ( 88 ). In addition, antioxidants such as NAC did not attenuate TGF- -mediated HO-1 induction in these cells, suggesting that the induction was potentially independent of oxidant stress ( 88 ). Further studies to examine the cellular consequences and molecular mechanisms of HO-1 gene expression in response to TGF- will be important in the pathogenesis of renal injury.4 O3 Q) @0 k. X% c* J; p
/ Y  `% `) o- e! P) c2 o
Other growth factors, such as PDGF ( 60 ), hepatocyte growth factor (HGF) ( 211 ), VEGF ( 65 ), and nerve growth factor (NGF) ( 192 ), have also been reported to induce HO-1 gene expression in nonrenal cells. Studies with HGF, a growth factor that promotes regeneration of renal epithelial cells ( 162 ), have suggested that hypoxia-inducible factor-1 (HIF-1 ) may regulate HO-1 gene expression in HepG2 cells ( 211 ). VEGF upregulates HO-1 protein expression in vivo in chick embryo chorioallantoic membranes by a mechanism dependent on an increase in cytosolic calcium levels and activation of protein kinase C ( 65 ). NGF-mediated HO-1 induction by a phosphatidylinositol 3-kinase/Akt-dependent pathway has recently been reported in dopaminergic PC12 cells ( 192 ).
  |7 Z! O8 V: W0 d+ @
* F: n4 h2 A" G& c, `. r' J% xAngiotensin II0 c+ H5 e- A( s3 w

# u+ s  z# r& k* c7 YAngiotensin II is an important mediator involved in the pathophysiology of renal injury. In addition to the well-known hemodynamic effects ( 6 ), angiotensin II also has nonhemodynamic effects, including the activation of several growth factors, cytokines, and the generation of reactive oxygen species in the vasculature and the kidney ( 6, 32, 83, 138, 166, 173 ). Angiotensin II is a potent inducer of HO-1, both in vivo in the intact rat kidney and in vitro in LLC-PK 1, rat renal proximal tubular epithelial, and mesangial cells ( 6, 29, 83, 97 - 99 ). However, the molecular regulation of HO-1 by angiotensin II has not as yet been reported. Systemic administration of angiotensin II using miniosmotic pumps resulted in significant upregulation of HO-1 in rat renal proximal tubules ( 83, 98 ). The induction of HO-1 by angiotensin II is not directly related to elevated blood pressure, because HO-1 was not induced in a rat model of hypertension after DOCA-salt ( 83 ) or norepinephrine infusion ( 6 ). HO-1 activation has also been reported in the rat heart ( 97 ) and aorta ( 100 ) after systemic angiotensin II administration.
: ?& f' ^- P1 \. W5 \" b/ C4 I  G5 G' K- k- y
Increased HO activity, with chemical inducers such as hemin, reverses angiotensin II-mediated decreases in glomerular filtration rate and increases in proteinuria, whereas HO inhibition leads to worsening of glomerular filtration rate and proteinuria ( 6 ), suggesting that HO-1 may provide a protective role to the potential injurious effects of angiotensin II in the kidney. Recent studies have shown that overexpression of HO-1 using a retroviral vector in human endothelial cells or hemin pretreatment in rat proximal tubular cells decreases angiotensin II-mediated cell injury ( 138 ). In our limited studies, we have not observed HO-1 induction in human proximal tubular or aortic endothelial cells using angiotensin II or angiotensin IV, its breakdown product (Hill-Kapturczak N and Agarwal A, unpublished observations). It is possible that interspecies variations may account for the differences observed in the human vs. other rodent or porcine cell lines.% R- u  O$ J( @1 B
" B: O& t& H2 X# T# J* B
Hypoxia+ S* k' P# W2 L+ N# O

5 @. a8 R! G. C* e  ^/ l- ~Another significant modulator of HO-1 gene expression is hypoxia. Interestingly, hypoxia induces HO-1 in rodent, bovine, and monkey cells but represses HO-1 expression in three different human cell lines (A549 human lung cancer cells, human umbilical vein endothelial cells, and human glioblastoma cells) ( 102, 113, 120, 146, 216, 240 ), indicating species-specific HO-1 regulation. The hypoxia-mediated repression of the human HO-1 gene is associated with activation of Bach1. The MARE located immediately downstream of the CdRE at approximately -4.0 kb of the human HO-1 promoter is required for hypoxia-mediated repression of HO-1 via Bach1 ( 113 ).0 Q  \; ?; S/ O3 }' O9 _7 Q
. I; E. D- F( d" D% D7 Q
Hypoxia significantly increases HO-1 mRNA expression by transcriptional activation in rat renal medullary interstitial cells ( 240 ). Using inhibitors of HIF-1 degradation and cis -element oligonucleotide decoys to block HIF-1, HO-1 induction was attenuated, suggesting a role for HIF-1 in hypoxia-mediated HO-1 activation ( 240 ). HIF-1 is a basic helix-loop-helix transcription factor whose expression is inducible by lower than normal levels of tissue oxygen. It acts by dimerizing and binding to a hypoxia-responsive element (HRE) in the promoter of hypoxia-responsive genes. Lee et al. ( 120 ) have demonstrated that HIF-1 mediates transcriptional activation of the mouse HO-1 gene in response to hypoxia and that the region responsible for this is located at approximately -9.5 kb upstream of the transcriptional start. Recent studies have reported the upregulation of HIF-1 in the border zone of rat renal segmental infarcts in tubular and glomerular cells, capillary endothelial cells, and infiltrating macrophages ( 190 ). HIF-1 expression colocalized with hypoxia-inducible genes, including HO-1 and VEGF. However, Wood et al. ( 236 ), using a mutant Chinese hamster ovary cell line for HIF-1, demonstrated that HO-1 expression was largely independent of hypoxia and did not require a functional HIF-1. Recent studies have demonstrated that the MAPK pathway, particularly p38 MAPK and MEK1, mediates the signal transduction pathway for hypoxia-mediated HO-1 induction in rodent cells ( 102 ). AP-1 DNA-binding activity is also increased and appears to be involved in HO-1 induction ( 191 ). Studies have demonstrated that HIF-1 regulates hypoxia-mediated HO-1 induction in rat aortic vascular smooth muscle cells, but not in rat pulmonary artery endothelial cells, where AP-1 is required ( 80 ). These data suggest that the regulation of the HO-1 gene in response to hypoxia in the pulmonary vasculature differs from the systemic circulation.8 N% X; \+ w) N1 e5 _; v
1 U" U" n; X( ~
Osmotic Stress, F3 ?- u. I3 k& r, \: i" F/ ^

7 |$ ]* R  o3 kRecent studies have suggested that the HO-1 induction may represent an adaptive response to changes in tonicity in the kidney ( 226, 255 ). Tian et al. ( 226 ) have demonstrated that urea and hypertonicity induce HO-1 gene expression in murine renal medullary cells via transcriptional activation. A 4.5-kb human HO-1 promoter was modestly activated in response to urea, but deletion of the CdRE did not affect promoter activity, suggesting that urea-mediated HO-1 induction was not dependent on the CdRE in these cells ( 226 ). Antioxidants such as NAC blocked urea-mediated HO-1 induction. In human hepatoma HepG2 cells, changes in osmolality by sugar molecules during glucose deprivation does not induce HO-1, suggesting a differential regulation of the mouse and human HO-1 genes in response to osmotic stress ( 40 ). It is interesting to note that hyperosmolarity suppresses HO-1 induction in response to hemin and pretreatment with the organic osmolyte betaine restored hemin-mediated HO-1 induction in rat hepatocytes ( 127 ).3 \/ _" C" H$ v! I: f

9 _" h- g/ v9 i$ F9 ~Cytokines and Endotoxin
5 j% Q! w6 L' i/ e4 B- Q! T! `/ [+ H9 b: I1 U7 z
Several inflammatory conditions such as atherosclerosis, transplant rejection, acute glomerulonephritis, and sepsis are associated with marked induction of HO-1 ( 3, 77, 229, 233 ). More importantly, the induction of HO-1 is a protective response in these disease settings ( 77, 96, 108, 121, 203 ). Proinflammatory cytokines and LPS, important mediators of these disorders, have been shown to activate HO-1, and it has been suggested that such induction occurs as an adaptive response to inflammatory stress ( 35, 37, 115, 175, 189, 241 ). However, most of these studies have been reported in rodent cells with the exception of a few in human cells ( 221, 222 ). IL-1 and TNF- transcriptionally activate HO-1 mRNA in human umbilical vein endothelial cells ( 221 ). NAC, an antioxidant, and mepacrine, a phospholipase A 2 inhibitor, blocked cytokine-mediated HO-1 induction. In addition, curcumin, an AP-1 inhibitor, decreased cytokine induction of HO-1 mRNA ( 222 ). However, curcumin by itself has been reported to induce HO-1 ( 26 ).2 {7 F5 [1 t0 a3 B2 }5 o3 \1 l

: u8 r. L1 k* X/ i+ aIL-1 has also been shown to induce HO-1 mRNA in rat mesangial cells ( 223 ). LPS also induces HO-1 via generation of hydrogen peroxide, and the transcription factor NF-kB has been implicated in such induction in mouse M1 myeloleukemia cells ( 115 ). The combination of LPS and interferon- induces HO-1 in murine mesangial cells ( 223 ). Some reports have demonstrated that LPS-mediated HO-1 induction occurs through IL-1, whereas others have shown that IL-1 knockout mice still respond to LPS and that the cytokine TNF- is actually responsible for HO-1 upregulation ( 170 ). IFN- with LPS or TNF- induces HO-1 in human monocytic cells. IL-6, an important mediator involved in the regulation of the acute-phase response to injury, induces HO-1 in Hep3B and HepG2 cells but not in human umbilical vein endothelial cells and murine macrophages ( 73, 141 ). IL-11, a member of the IL-6 family of cytokines, has also been reported to induce HO-1 in HepG2 cells ( 72 ). We have tested the effects of cytokines alone and in combination and found no induction of HO-1 in human renal epithelial cells ( Fig. 5 ), results consistent with the findings of Takahashi et al. ( 213 ), who reported no induction with IFN- and IL-1 in human glioblastoma cells, where in fact these cytokines actually repress HO-1 activation by SNP, cadmium, and hemin. The reasons for the differential responses in the human and rodent species are not entirely clear but may be similar to the interspecies differences observed in the regulation of the iNOS gene ( 140 ).9 N0 b% q7 A4 S8 `5 _
+ W: X2 o' Y6 j: E; [, _
Anti-inflammatory cytokines also activate HO-1 gene expression. For instance, IL-10 induces HO-1 in murine macrophages ( 121 ). More importantly, the protective effect of IL-10 in a murine model of LPS-induced septic shock was significantly attenuated by inhibition of HO enzyme activity, suggesting that the anti-inflammatory effects of IL-10 were mediated via HO-1 induction ( 121 ). Recent findings have demonstrated the efficacy of IL-10 in animal models of acute renal injury secondary to ischemia-reperfusion and cisplatin ( 50 ), models wherein HO-1 has also been shown to be protective ( 1, 200, 201, 234 ). Given the role of HO-1 induction in mediating the protective effects of IL-10 in sepsis ( 121 ), it is tempting to speculate that the effect of IL-10 in acute renal injury also occurs via HO-1 activation. IL-13, an immunoregulatory cytokine that is a key mediator in allergic inflammation, has also been shown to induce HO-1 ( 110 ). Similar to the effects of IL-10 in sepsis, HO-1 induction has been suggested to mediate the effects of IL-13 in vivo in rat cardiac allografts and in vitro in human umbilical vein endothelial cells ( 110 ).3 R) b( `* |% L6 B% G7 x

+ j2 [. k. j6 V' P* Q, _SUMMARY
# D1 T1 S* H/ u% d2 v) M. z) p8 c7 R7 j" O
Induction of HO-1 plays an important role in the pathophysiology of several diseases involving the kidney as well as other organ systems. The biological implications of the HO-1-catalyzed reaction have gone far beyond the initial description of its function as the rate-limiting enzyme in heme degradation in reticuloendothelial cells, where increased levels of heme are derived from hemoglobin, released from red blood cells completing their life cycle of 120 days. The reaction products of HO-1 have both beneficial and potentially injurious effects. There is considerable heterogeneity in the tissue response to injury, and induction of HO-1 may not always be beneficial. An optimal level of HO-1 induction to provide a therapeutic level of a "reaction product" may be necessary and would depend on the setting. Several features suggest differences in the regulation of the mouse and human HO-1 gene. While the mouse HO-1 gene has been well characterized and studied, the human HO-1 gene requires further characterization. Experiments using additional promoter constructs, chromatin structure analysis, and in vivo footprinting are currently underway in our laboratory to delineate the region(s) of the human HO-1 gene that control induction. It is anticipated that the knowledge gained by studies involving the molecular regulation of human HO-1 gene expression will allow for the fine-tuning of HO-1 gene expression in disease states and hence the ability to exploit the cytoprotective effects of HO-1. Strategies to target and achieve regulated expression of HO-1 will have significant therapeutic implications in several clinical settings involving the kidney.7 B* A  d) j9 }& a

( J7 P$ b7 Y, h; LGRANTS" _3 j3 ~+ J5 s8 B. Q8 H, F

4 P. B  w. ]7 |# n' s2 H. QThis work was supported by National Institutes of Health Grants R01-DK-59600 and R01-HL-68157 (to A. Agarwal) and K01-DK-02902 (to N. Hill-Kapturczak). E. Sikorski is supported by a National Kidney Foundation fellowship award., }( K, J& i$ T

! r4 b$ ~/ D* ~  C4 n7 vAddress for reprint requests and other correspondence: A. Agarwal, Dept. of Medicine, Div. of Nephrology, ZRB 614, 1530 3rd. Ave. South, Univ. of Alabama at Birmingham, Birmingham, AL 35294 (E-mail: agarwal{at}uab.edu
2 @+ T' m6 ]) M# A# H: y3 o3 O1 p          【参考文献】* ]( e' f" X5 D5 K* S1 x3 P
Agarwal A, Balla J, Alam J, Croatt AJ, and Nath KA. Induction of heme oxygenase in toxic renal injury: a protective role in cisplatin nephrotoxicity in the rat. Kidney Int 48: 1298-1307, 1995.
/ X3 q6 p2 x0 j  s0 y
( t& M4 x# F( |# g. @1 T/ j* w5 |1 l$ F. o. v2 p+ E

" K1 ?0 x& }% p$ A( vAgarwal A, Balla J, Balla G, Croatt AJ, Vercellotti GM, and Nath KA. Renal tubular epithelial cells mimic endothelial cells upon exposure to oxidized LDL. Am J Physiol Renal Fluid Electrolyte Physiol 271: F814-F823, 1996.
( }! I' C% B" j% j- U3 \
9 o' V6 [* P- E% C4 \3 B
$ X6 `, X4 \8 c* O+ C# w# Y- Z7 p9 Q! G/ }" E
Agarwal A, Kim Y, Matas AJ, Alam J, and Nath KA. Gas-generating systems in acute renal allograft rejection in the rat. Co-induction of heme oxygenase and nitric oxide synthase. Transplantation 61: 93-98, 1996.
4 V$ }# b; B9 r4 ~* V
0 ?, Y4 c+ v+ ?3 g# i9 _* I. z
/ f" A( |2 y2 _5 z& p; P0 B  Z; y& [  U8 C
Agarwal A and Nick HS. Renal response to tissue injury: lessons from heme oxygenase-1 gene ablation and expression. J Am Soc Nephrol 11: 965-973, 2000.. b& i$ P! }- v6 R) @
/ j  y, o' K- ]! g' L
9 P0 Z7 [5 b  O. _& v- C& p
  z" t1 j1 N$ R! P
Agarwal A, Shiraishi F, Visner GA, and Nick HS. Linoleyl hydroperoxide transcriptionally upregulates heme oxygenase-1 gene expression in human renal epithelial and aortic endothelial cells. J Am Soc Nephrol 9: 1990-1997, 1998.
8 W) f. }* V. a) Z
+ d& N% U! ^$ w" s8 H9 z6 ], n. ]! v7 X' y) A8 k$ a! l
" O$ g& d( i' A6 t% g5 x6 G
Aizawa T, Ishizaka N, Taguchi J, Nagai R, Mori I, Tang SS, Ingelfinger JR, and Ohno M. Heme oxygenase-1 is upregulated in the kidney of angiotensin II-induced hypertensive rats: possible role in renoprotection. Hypertension 35: 800-806, 2000.* `# }* V7 q7 _! e# Q& I

" `( T* m& X) j! B( Y
+ H; J. E6 i& v9 I
8 R5 D3 m: N1 B0 E$ JAlam J. Multiple elements within the 5' distal enhancer of the mouse heme oxygenase-1 gene mediate induction by heavy metals. J Biol Chem 269: 25049-25056, 1994.2 p1 }* t% s, n$ K+ A9 U# S# @7 m

9 I( B1 D) d- U7 {9 T) {) `2 o' x( g- s; M0 B

, Q6 L- s/ z: e8 g9 JAlam J, Cai J, and Smith A. Isolation and characterization of the mouse heme oxygenase-1 gene. Distal 5' sequences are required for induction by heme or heavy metals. J Biol Chem 269: 1001-1009, 1994.% ?' _- L) l6 ?$ c- C" r, A
7 J# v+ P$ ^% L) q4 g+ p3 l" A
6 W: y# p. w+ y6 G
' ~' o# ^, g9 X: h8 C# A! V
Alam J, Camhi S, and Choi AM. Identification of a second region upstream of the mouse heme oxygenase-1 gene that functions as a basal level and inducer-dependent transcription enhancer. J Biol Chem 270: 11977-11984, 1995.( G3 l: m/ l5 C- Q7 d, a
2 t' Z! M" I. H( f3 V
; F' j# V, O5 l- {9 p5 z
3 e8 v* ^; n$ y8 M1 l
Alam J and Den Z. Distal AP-1 binding sites mediate basal level enhancement and TPA induction of the mouse heme oxygenase-1 gene. J Biol Chem 267: 21894-21900, 1992.5 [6 a4 r% O# _( T% _

7 e. g( M; V" B6 P5 ~
1 \  h  @( a1 P% b( H) z- @  B, T& g+ e2 ?3 F8 l
Alam J, Killeen E, Gong P, Naquin R, Hu B, Stewart D, Ingelfinger JR, and Nath KA. Heme activates the heme oxygenase-1 gene in renal epithelial cells by stabilizing Nrf2. Am J Physiol Renal Physiol 284: F743-F752, 2003., g  {# q& N; C: @
! Z: a0 h$ }  D3 N3 i3 L

1 ^, A" ]7 B% O) d$ I8 ~7 ^, l# A4 C4 L" O: I
Alam J, Shibahara S, and Smith A. Transcriptional activation of the heme oxygenase gene by heme and cadmium in mouse hepatoma cells. J Biol Chem 264: 6371-6375, 1989.
, P" C( ^' @3 H$ `! W! c
* i) r9 i" Z9 B$ h
# Q$ Z4 O; e, e* ]2 s
8 ~) a( J: T6 {' y( e9 R) ~0 @) a5 V, l; TAlam J and Smith A. Receptor-mediated transport of heme by hemopexin regulates gene expression in mammalian cells. J Biol Chem 264: 17637-17640, 1989.
1 P1 u4 p# L% e, O5 o5 v5 ]! k
5 X* I( e. g. R: Q# b2 N
1 n$ `) ?6 p. j8 v: L4 O: O, H* a: _
Alam J, Stewart D, Touchard C, Boinapally S, Choi AM, and Cook JL. Nrf2, a Cap'n'Collar transcription factor, regulates induction of the heme oxygenase-1 gene. J Biol Chem 274: 26071-26078, 1999.
4 U2 N3 B0 x/ q  S4 `7 w# r8 ]
5 `% q# y  J8 J2 H4 j* t
2 ?' y% s1 g, X3 }9 A# e9 L7 n9 c$ }9 r& V: c; w
Alam J, Wicks C, Stewart D, Gong P, Touchard C, Otterbein S, Choi AM, Burow ME, and Tou J. Mechanism of heme oxygenase-1 gene activation by cadmium in MCF-7 mammary epithelial cells. Role of p38 kinase and Nrf2 transcription factor. J Biol Chem 275: 27694-27702, 2000.8 F1 E5 o# ?% T. @  `7 ^/ s- {) y
* w8 \+ K4 C( t( X* S5 T

+ z/ E* P6 ?' r: s2 N3 ~* k- z& ?. ^
  ~' J, }. @0 C  J' B0 x; y4 `: {Alcaraz MJ, Habib A, Creminon C, Vicente AM, Lebret M, Levy-Toledano S, and Maclouf J. Heme oxygenase-1 induction by nitric oxide in RAW 264. 7 macrophages is upregulated by a cyclo-oxygenase-2 inhibitor. Biochim Biophys Acta 1526: 13-16, 2001.# {# G2 S6 J) T$ B8 [

' p& ?9 v$ n0 D# C/ b. u5 d5 o) E4 E. h2 L' t% A
/ d1 o9 L; v! A4 t9 t, @8 r* \
Amersi F, Buelow R, Kato H, Ke B, Coito AJ, Shen XD, Zhao D, Zaky J, Melinek J, Lassman CR, Kolls JK, Alam J, Ritter T, Volk HD, Farmer DG, Ghobrial RM, Busuttil RW, and Kupiec-Weglinski JW. Upregulation of heme oxygenase-1 protects genetically fat Zucker rat livers from ischemia/reperfusion injury. J Clin Invest 104: 1631-1639, 1999.; J4 D9 {% }. X4 E9 G% l4 j. E

+ S  C3 ~8 \+ x. \* n2 s. V# N) V+ B- h; t

' Q) M9 D5 }# f4 V6 PAmersi F, Shen XD, Anselmo D, Melinek J, Iyer S, Southard DJ, Katori M, Volk HD, Busuttil RW, Buelow R, and Kupiec-Weglinski JW. Ex vivo exposure to carbon monoxide prevents hepatic ischemia/reperfusion injury through p38 MAP kinase pathway. Hepatology 35: 815-823, 2002.
$ f8 h$ i4 [$ a3 F+ b2 j" e# S; a
% B: n6 r, n1 E  }8 Q
) \1 G7 q2 ]5 f6 N0 o% b
Applegate LA, Luscher P, and Tyrrell RM. Induction of heme oxygenase: a general response to oxidant stress in cultured mammalian cells. Cancer Res 51: 974-978, 1991.! g+ O' s" r8 T- y2 \

7 X+ G3 C8 K* W8 s
0 \7 b2 h; m& ?7 T2 @( C' r1 Z2 N2 {! I& V8 u
Atfi A, Buisine M, Mazars A, and Gespach C. Induction of apoptosis by DPC4, a transcriptional factor regulated by transforming growth factor- through stress-activated protein kinase/c-Jun N-terminal kinase (SAPK/JNK) signaling pathway. J Biol Chem 272: 24731-24734, 1997.7 ]7 J) a1 m! k  J6 K" i5 f
3 o3 O0 z2 \( G5 x! j2 l

: @2 Q' x, P, G4 }2 H/ k+ F( _, Q
$ b% Y3 Z0 Y" _% z( @Backenroth R, Schuger L, Wald H, and Popovtzer MM. Glycerol-induced acute renal failure attenuates subsequent HgCl 2 -associated nephrotoxicity: correlation of renal function and morphology. Ren Fail 20: 15-26, 1998.
; E4 `, d% T4 l+ _" {" S5 Q- i% g0 K$ g$ U2 |! l0 e) v
# d$ @2 ^7 u7 G. U4 M8 E0 ]& ^

; d1 B+ ]0 I" [& Y4 \/ xBalla G, Jacob HS, Balla J, Rosenberg M, Nath K, Apple F, Eaton JW, and Vercellotti GM. Ferritin: a cytoprotective antioxidant strategem of endothelium. J Biol Chem 267: 18148-18153, 1992.% \1 f: U4 X/ z
" \0 o: s' I' S; Q3 H( y

& G, X$ U! \! F$ C) \5 C, u: w4 I  K4 P  `0 x9 I  L' v; ~
Balla G, Vercellotti GM, Muller-Eberhard U, Eaton J, and Jacob HS. Exposure of endothelial cells to free heme potentiates damage mediated by granulocytes and toxic oxygen species. Lab Invest 64: 648-655, 1991.
% ~; c; Z( k9 [5 j$ G# s4 M! k3 s* a, Y! z) M4 e8 x
$ K" Z% y6 [. u+ B' i0 M. i

; @5 G5 v; r1 [8 t- j, j7 i$ XBalla J, Jacob HS, Balla G, Nath K, Eaton JW, and Vercellotti GM. Endothelial-cell heme uptake from heme proteins: induction of sensitization and desensitization to oxidant damage. Proc Natl Acad Sci USA 90: 9285-9289, 1993.
. N/ R8 n' M' T
3 |$ x4 X) }+ M# |- J: D
! q! V  T6 ^* E: C: |1 B
: X6 Y- A4 a; z: P/ y4 ZBalla J, Nath KA, Balla G, Juckett MB, Jacob HS, and Vercellotti GM. Endothelial cell heme oxygenase and ferritin induction in rat lung by hemoglobin in vivo. Am J Physiol Lung Cell Mol Physiol 268: L321-L327, 1995.7 Z5 R' b! n, l/ S8 F( e/ e/ c

, t  T$ c: I* N# J6 z3 ^& U8 V. {: e2 F2 y  M, c! F, e

( v. s  H3 O( R; Z+ R* G2 `& ZBalogun E, Hoque M, Gong P, Killeen E, Green CJ, Foresti R, Alam J, and Motterlini R. Curcumin activates the heme oxygenase-1 gene via regulation of Nrf2 and the antioxidant responsive element. Biochem J 371: 887-895, 2003.& f; [& N6 s/ B4 l. \# o' o7 n% a

! d" T7 h: o1 w9 j
! T, ~+ d7 h9 K( z  ^' i9 Y. Q+ c6 ]1 D$ \) U" Y, G
Baranano DE, Rao M, Ferris CD, and Snyder SH. Biliverdin reductase: a major physiologic cytoprotectant. Proc Natl Acad Sci USA 99: 16093-16098, 2002.
7 ?/ h  [6 O4 O4 l; ?  }: O5 V1 j1 I( t( B  O  t! o9 B
8 I) ]4 {8 N6 `8 p& t* j  b0 y
( c9 Z% v0 {: {3 J# w
Beckman JS, Beckman TW, Chen J, Marshall PA, and Freeman BA. Apparent hydroxyl radical production by peroxynitrite: implications for endothelial injury from nitric oxide and superoxide. Proc Natl Acad Sci USA 87: 1620-1624, 1990.
8 g% p# A- b- P9 r! S
0 U4 {& q% u6 I0 a7 d2 P$ C, ^2 w; s
# N2 K! j9 D$ m& F: H) Y
Bhaskaran M, Reddy K, Radhakrishnan N, Franki N, Ding G, and Singhal PC. Angiotensin II induces apoptosis in renal proximal tubular cells. Am J Physiol Renal Physiol 284: F955-F965, 2003.& x! w' I' A9 R, Y$ F

9 c4 b! y1 x/ E# X+ @$ D/ X- C0 x( _4 U2 o- w1 w

3 z" R% F# S' j& qBlank V and Andrews NC. The Maf transcription factors: regulators of differentiation. Trends Biochem Sci 22: 437-441, 1997.$ R) J% e% x- q  V0 O$ D; A* V

" g( G: ?# q& i5 R) X
2 L% I" H: `+ c% ?
, {2 N. |- P2 d! l# [( E2 kBlouin JL, Duriaux Sail G, Guipponi M, Rossier C, Pappasavas MP, and Antonarakis SE. Isolation of the human BACH1 transcription regulator gene, which maps to chromosome 21q22.1. Hum Genet 102: 282-288, 1998.
  C* o8 L' A5 ]0 c' \. @2 U1 {" g+ v  l% `) D

* ]: E: g/ m( y. G# ?
+ u0 {* _' P) F2 H9 ^& X. uBorder WA and Noble NA. Interactions of transforming growth factor- and angiotensin II in renal fibrosis. Hypertension 31: 181-188, 1998.% R' i. S& \  _  C  a
$ h. _$ \% S$ x# B) W( i% Z1 p

9 R. Z$ T' s4 I8 i6 p4 R% J
2 g5 c# Y( B" E9 VBouton C and Demple B. Nitric oxide-inducible expression of heme oxygenase-1 in human cells. Translation-independent stabilization of the mRNA and evidence for direct action of nitric oxide. J Biol Chem 275: 32688-32693, 2000./ |& {" U! E5 }# v% \. C  q9 I
# ]: y+ `& ]9 D6 o/ D% R
0 [1 m0 C& f4 [9 {3 p' Z9 [
: i1 L( C- C6 n9 i/ F$ x
Brouard S, Otterbein LE, Anrather J, Tobiasch E, Bach FH, Choi AM, and Soares MP. Carbon monoxide generated by heme oxygenase 1 suppresses endothelial cell apoptosis. J Exp Med 192: 1015-1026, 2000.
0 |0 v$ o) b/ ~" p3 z+ y4 u) D, |! z6 w5 {! K; O" }
  b. i) B; x" b$ B) s
* u& c/ P! `! \' A
Camhi SL, Alam J, Otterbein L, Sylvester SL, and Choi AM. Induction of heme oxygenase-1 gene expression by lipopolysaccharide is mediated by AP-1 activation. Am J Respir Cell Mol Biol 13: 387-398, 1995., ]; J/ O, y0 N  U
; k. _7 Z3 H" F
; N; Y) S! E) F6 T; W& s( u; Q8 N& f

8 u( @# [" j7 T0 {$ G0 O+ U$ YCamhi SL, Alam J, Wiegand GW, Chin BY, and Choi AM. Transcriptional activation of the HO-1 gene by lipopolysaccharide is mediated by 5' distal enhancers: role of reactive oxygen intermediates and AP-1. Am J Respir Cell Mol Biol 18: 226-234, 1998.
+ ~' s; v. P: Y6 B
) V& m4 m" W( e) a# R! |5 B9 K  U% h8 j0 l! K% S; ~2 |

0 n' n. v# |! y' p% TCantoni L, Rossi C, Rizzardini M, Gadina M, and Ghezzi P. Interleukin-1 and tumour necrosis factor induce hepatic haem oxygenase. Feedback regulation by glucocorticoids. Biochem J 279: 891-894, 1991.
" {6 n/ z" m+ e- n
' L' a. I7 {2 {& f$ f) V% m2 l4 b" ~) E

* i( k& }* S, G! i0 y- xCary SP and Marletta MA. The case of CO signaling: why the jury is still out. J Clin Invest 107: 1071-1073, 2001.
3 |* y% N6 ~$ ?! m9 X# f
! G6 }4 e+ a7 t% Q/ ~+ M' j2 T
% g$ a! U8 s- t
) I) W8 ]* y* QChan K, Lu R, Chang JC, and Kan YW. NRF2, a member of the NFE2 family of transcription factors, is not essential for murine erythropoiesis, growth, and development. Proc Natl Acad Sci USA 93: 13943-13948, 1996.
5 d% \: f* i8 X: j* [+ a1 c9 l2 ?) G. A5 ~' d
" [2 E" e: k9 P

' x# n+ d. V8 H/ K& MChang SH, Barbosa-Tessmann I, Chen C, Kilberg MS, and Agarwal A. Glucose deprivation induces heme oxygenase-1 gene expression by a pathway independent of the unfolded protein response. J Biol Chem 277: 1933-1940, 2002.* u* z! F6 s6 g7 l

8 p5 p. w8 p5 b- X' R& k" a7 p
5 y! R) O( c1 i5 t: R+ T: f. A
8 B" b( o  w/ O# |( m$ ^% OChen K and Maines MD. Nitric oxide induces heme oxygenase-1 via mitogen-activated protein kinases ERK and p38. Cell Mol Biol (Noisyle-grand) 46: 609-617, 2000.
5 h0 g0 H! i. |8 ~6 Z  S) i* _8 |3 w2 D. H2 V
- U# e* s9 N4 P) v

9 S$ B$ l; S& A3 U* Y$ m( PChen YH, Lin SJ, Lin MW, Tsai HL, Kuo SS, Chen JW, Charng MJ, Wu TC, Chen LC, Ding YA, Pan WH, Jou YS, and Chau LY. Microsatellite polymorphism in promoter of heme oxygenase-1 gene is associated with susceptibility to coronary artery disease in type 2 diabetic patients. Hum Genet 111: 1-8, 2002.
% V, G8 Z- }! I, [4 h8 t( R3 P) B6 d/ C, P# N) n  }7 C& |& y3 P- \$ Z+ }5 v/ v

/ p0 i& v8 J) K3 T1 C9 m5 l# i# ^7 o& l# A$ t3 e( i- m0 m6 ?
Cho HY, Jedlicka AE, Reddy SP, Kensler TW, Yamamoto M, Zhang LY, and Kleeberger SR. Role of NRF2 in protection against hyperoxic lung injury in mice. Am J Respir Cell Mol Biol 26: 175-182, 2002.. n& V4 R6 J: E2 I/ y

4 b' P; }0 Z6 o  a0 E( c2 x  C) d  g9 y: D

, O/ p9 ~7 e$ D# _Choi AM and Alam J. Heme oxygenase-1: function, regulation, and implication of a novel stress-inducible protein in oxidant-induced lung injury. Am J Respir Cell Mol Biol 15: 9-19, 1996.% ~; |7 x% N  z% b( E

2 Z5 N$ P! t; U: V7 [9 U# t( T6 }2 C* X9 F9 }& f, M: g: L

4 e' y5 \$ X7 B/ v; pClark JE, Foresti R, Sarathchandra P, Kaur H, Green CJ, and Motterlini R. Heme oxygenase-1-derived bilirubin ameliorates postischemic myocardial dysfunction. Am J Physiol Heart Circ Physiol 278: H643-H651, 2000.
. W4 g$ b/ f" }' X- @5 T
( v2 S* E9 u* J  S+ @& F% U& Z3 w; j7 l) J3 w; L& q) X
. k: E4 j7 l3 U! q
Cuturi MC, Christoph F, Woo J, Iyer S, Brouard S, Heslan JM, Pignon P, Soulillou JP, and Buelow R. RDP1258, a new rationally designed immunosuppressive peptide, prolongs allograft survival in rats: analysis of its mechanism of action. Mol Med 5: 820-832, 1999.
+ b% o4 C) z6 K( {/ d0 x7 x  `7 E9 B- T/ T6 I1 j) Y6 I8 p

2 t8 }6 }" m  q& K3 [+ n9 c( L0 C" {5 h
Dai C, Yang J, and Liu Y. Transforming growth factor- 1 potentiates renal tubular epithelial cell death by a mechanism independent of Smad signaling. J Biol Chem 278: 12537-12545, 2003.; d: d( w% C2 N6 Y2 A( n$ \

6 h( g3 ~3 p: E$ J( L" d# @7 N; [. G1 m& g4 K

0 J" j" o$ s) S3 SDatta PK, Koukouritaki SB, Hopp KA, and Lianos EA. Heme oxygenase-1 induction attenuates inducible nitric oxide synthase expression and proteinuria in glomerulonephritis. J Am Soc Nephrol 10: 2540-2550, 1999./ J" F. V9 C; l+ W3 i1 D  T' J  T4 |
# e( J( M4 P0 T% W5 F3 J
# z$ T  H9 i2 b0 R& z  H7 u

* X) C: ?' r2 }! m* [; S5 D& CDatta PK and Lianos EA. Nitric oxide induces heme oxygenase-1 gene expression in mesangial cells. Kidney Int 55: 1734-1739, 1999.+ P0 S4 Q. E% L! K6 d
4 j% H( h+ M+ y7 y" {8 k- v

$ C7 s; [: X7 m: |2 |4 T1 T6 o  s$ Z) \& `
Deng J, Kohda Y, Chiao H, Wang Y, Hu X, Hewitt SM, Miyaji T, McLeroy P, Nibhanupudy B, Li S, and Star RA. Interleukin-10 inhibits ischemic and cisplatin-induced acute renal injury. Kidney Int 60: 2118-2128, 2001.
; T. Q- O8 i9 _2 g7 c6 K6 ^, C* ?1 ]
% H* s; i+ L' k: [9 R4 k
& C1 e7 p" d' x2 x' W( F9 k$ J2 `
Dennery PA. Regulation and role of heme oxygenase in oxidative injury. Curr Top Cell Regul 36: 181-199, 2000., @! a+ [; l* J6 {3 A
' ^: a7 c0 R, G# Z& d: d  h; C
+ ^# p- H4 q2 P: U) @
0 X% f) P" `1 O) g8 [: G6 x% _
Dennery PA, Seidman DS, and Stevenson DK. Neonatal hyperbilirubinemia. N Engl J Med 344: 581-590, 2001.
% Q% U+ o& V$ Q3 z% \8 D# i" V
5 @6 _, X2 r6 K4 t, H9 _: [
+ I' R5 R1 k' @% |% _- H2 r/ y4 w/ X( P
Dennler S, Itoh S, Vivien D, ten Dijke P, Huet S, and Gauthier JM. Direct binding of Smad3 and Smad4 to critical TGF- -inducible elements in the promoter of human plasminogen activator inhibitor-type 1 gene. Embo J 17: 3091-3100, 1998.
) r$ ~0 T7 w; R2 n5 x2 S( z$ A  z

0 H0 R% {8 F2 e/ l: Q- C) ?, s$ K+ Z1 z( ~2 r7 L/ C. b; O7 J
Deramaudt BM, Remy P, and Abraham NG. Upregulation of human heme oxygenase gene expression by Ets-family proteins. J Cell Biochem 72: 311-321, 1999. <a href="/cgi/external_ref?access_num=10.1002/(SICI)1097-4644(19990301)72:3
0 a9 E0 K! Z8 Q7 m% Q2 w0 j" T  b" X8 v
3 ]( K! n2 b8 a
7 B* K. d0 |  g& Y
Dong Z, Lavrovsky Y, Venkatachalam MA, and Roy AK. Heme oxygenase-1 in tissue pathology: the Yin and Yang. Am J Pathol 156: 1485-1488, 2000.
' {  S# A8 ~/ B/ `3 q$ t
' `+ z. H/ @8 b* \
- ]: ?' ^8 f$ Q
; s5 {+ _: P0 L% VDore S, Takahashi M, Ferris CD, Zakhary R, Hester LD, Guastella D, and Snyder SH. Bilirubin, formed by activation of heme oxygenase-2, protects neurons against oxidative stress injury. Proc Natl Acad Sci USA 96: 2445-2450, 1999.
1 W* y3 B1 Q4 B9 f2 V. A5 K
5 z. X& y8 k+ B0 _8 K* A* H" D( u. K- K; K4 h. Y( i

# u0 z2 y6 j7 A% tDuckers HJ, Boehm M, True AL, Yet SF, San H, Park JL, Clinton Webb R, Lee ME, Nabel GJ, and Nabel EG. Heme oxygenase-1 protects against vascular constriction and proliferation. Nat Med 7: 693-698, 2001.1 N1 S# y) A; |( f
& S) c* W6 C( S# \4 A/ z7 M
3 t- o( K" `$ a+ J: K5 _& v4 L
& `' f- O, ]+ L+ J7 r2 _
Durante W. Heme oxygenase-1 in growth control and its clinical application to vascular disease. J Cell Physiol 195: 373-382, 2003.2 _0 o" h! l; Z# q. {% A4 ^3 p# z

* w7 `5 @- a: |/ R  n+ b, q- H$ i. o9 q3 V
" H3 ?2 y) v4 v. k
Durante W, Kroll MH, Christodoulides N, Peyton KJ, and Schafer AI. Nitric oxide induces heme oxygenase-1 gene expression and carbon monoxide production in vascular smooth muscle cells. Circ Res 80: 557-564, 1997.* O# @& ?' c  R( _* o! q2 h
* J4 e( ], h6 P% g# l

- I+ ~) R) j5 U" B. V
; g* F3 H+ v: tDurante W, Peyton KJ, and Schafer AI. Platelet-derived growth factor stimulates heme oxygenase-1 gene expression and carbon monoxide production in vascular smooth muscle cells. Arterioscler Thromb Vasc Biol 19: 2666-2672, 1999.
3 o- @- ?* X  h0 N: |! s1 t- C5 t, n  C
. x( R6 ?: \/ i7 W7 ~! K

$ n2 U) l+ M& h% Y+ y. v0 S; [Durante W and Schafer AI. Carbon monoxide and vascular cell function. Int J Mol Med 2: 255-262, 1998.
" Y6 X. z4 y& I& D- ]* A# ^3 J, R2 k. m3 T# {: B! {- y

% H& b9 d) ?8 ?! ^0 w! k* h; K4 H9 H1 [" A
Engel ME, Datta PK, and Moses HL. Signal transduction by transforming growth factor- : a cooperative paradigm with extensive negative regulation. J Cell Biochem Suppl 30-31: 111-122, 1998.
8 d+ R! f, g# q* C, h
3 N1 m. a& D1 C
5 X7 _0 N; p8 k, t3 V
# G: h# W( _' S5 N/ X; v" R$ G$ dExner M, Schillinger M, Minar E, Mlekusch W, Schlerka G, Haumer M, Mannhalter C, and Wagner O. Heme oxygenase-1 gene promoter microsatellite polymorphism is associated with restenosis after percutaneous transluminal angioplasty. J Endovasc Ther 8: 433-440, 2001. <a href="/cgi/external_ref?access_num=10.1583/1545-1550(2001)008
* o% f# W! N: X) O0 F  h# B% G. b4 N. U; f7 d  g  k

' _/ z/ Z) i/ D3 j0 j5 d
- |" _+ x8 D, p& fFavatier F and Polla BS. Tobacco-smoke-inducible human haem oxygenase-1 gene expression: role of distinct transcription factors and reactive oxygen intermediates. Biochem J 353: 475-482, 2001.
- ^8 {' k+ w& i! P4 V9 ?) Q) X& V0 @4 p* B: a, ?& _% {" F
$ q! j, E) V' y/ e# f: y1 v8 `9 \
# f, G0 |8 l: X, ?
Fernandez M and Bonkovsky HL. Vascular endothelial growth factor increases heme oxygenase-1 protein expression in the chick embryo chorioallantoic membrane. Br J Pharmacol 139: 634-640, 2003.
# l4 n3 V3 v2 g- ^+ n2 v- o! K$ [5 p# `
8 ~' Z- v; Q. l/ H% n  {) s
. z. \2 Q+ t4 {8 c, ]2 a. T, D/ j
Ferris CD, Jaffrey SR, Sawa A, Takahashi M, Brady SD, Barrow RK, Tysoe SA, Wolosker H, Baranano DE, Dore S, Poss KD, and Snyder SH. Haem oxygenase-1 prevents cell death by regulating cellular iron. Nat Cell Biol 1: 152-157, 1999.
& V  \, S6 S5 ^+ X; j
9 t' c$ g& p9 ~7 E3 O( }( F9 _
0 Y- K9 ^& ^8 J3 Q# ?
' E& b9 Y$ F  T( D! H. t8 PFitzhugh AL and Keefer LK. Diazeniumdiolates: pro- and antioxidant applications of the "NONOates." Free Radic Biol Med 28: 1463-1469, 2000.4 m( K1 B6 ~) y& U9 y7 n) ?' k
8 B& G; l/ S3 v8 y4 ?; ~
6 {* ^! K  o, l) s# G

: G6 U( \* j  E+ ~9 e7 @) EFogg S, Agarwal A, Nick HS, and Visner GA. Iron regulates hyperoxia-dependent human heme oxygenase 1 gene expression in pulmonary endothelial cells. Am J Respir Cell Mol Biol 20: 797-804, 1999.
* Z/ o8 \) j( y4 R4 @, D
% x  O  z% U2 \* E% s( Y2 }
) z6 M: o% ?- a5 x! N+ K
- N5 ^& h8 [9 j) Y# ^1 ^1 @Foresti R, Clark JE, Green CJ, and Motterlini R. Thiol compounds interact with nitric oxide in regulating heme oxygenase-1 induction in endothelial cells. Involvement of superoxide and peroxynitrite anions. J Biol Chem 272: 18411-18417, 1997.1 J( Y+ C& A. \* Y4 G, T  k

  n5 h. ^, E8 }
5 ^, v& s/ ]  U# ]) d' ]3 \9 L, e8 ?! s# N' a3 _7 Q3 K, K
Foresti R, Sarathchandra P, Clark JE, Green CJ, and Motterlini R. Peroxynitrite induces haem oxygenase-1 in vascular endothelial cells: a link to apoptosis. Biochem J 339: 729-736, 1999.
3 [" _- z  r' \! q" w! x4 K, j: w. O# e, i
- B; s! u8 ]9 u' O( P/ i
" i( P" _# w# ?
Fujita T, Toda K, Karimova A, Yan SF, Naka Y, Yet SF, and Pinsky DJ. Paradoxical rescue from ischemic lung injury by inhaled carbon monoxide driven by derepression of fibrinolysis. Nat Med 7: 598-604, 2001.
$ Y% t0 v0 [/ Z9 D% I' X2 Z
9 k0 D8 E+ ]( _' a- p, f/ b
3 g: s8 z4 |# x$ }0 @3 n4 o! M5 k8 I( R, ~( o6 h
Fukuda Y and Sassa S. Effect of interleukin-11 on the levels of mRNAs encoding heme oxygenase and haptoglobin in human HepG2 hepatoma cells. Biochem Biophys Res Commun 193: 297-302, 1993., r; o/ G( A1 |2 |9 l2 t: e0 k
. e- Y. x8 D5 Y& p+ s$ y: }) U
  L, S7 ~; d3 l

+ r* R2 {$ ~7 [- l$ mFukuda Y and Sassa S. Suppression of cytochrome P450IA1 by interleukin-6 in human HepG2 hepatoma cells. Biochem Pharmacol 47: 1187-1195, 1994.
7 L* E: y# H% H" g2 E  d2 P) u3 \& T, E( T' Z, t& y
) `) g& k0 ^" z( i" j  l+ |
+ x  n1 C, G5 s' q
Galbraith R. Heme oxygenase: who needs it? Proc Soc Exp Biol Med 222: 299-305, 1999.4 I' t4 [4 F+ x3 W. S: v( u+ T
% O' H' c. G" G, W0 Q' C

+ ~& r& C0 x) [& _" m+ A! m
6 |! A/ k, }( f" [* I6 q1 ZGong P, Hu B, Stewart D, Ellerbe M, Figueroa YG, Blank V, Beckman BS, and Alam J. Cobalt induces heme oxygenase-1 expression by a hypoxia-inducible factor-independent mechanism in Chinese hamster ovary cells: regulation by Nrf2 and MafG transcription factors. J Biol Chem 276: 27018-27025, 2001.
; n7 k8 {3 |% y6 u/ y
2 y, X' m" i$ R
4 i% p# \& a. ~; g4 `
& b. d& p8 u5 Z$ V& @0 w( p1 @. M% XGong P, Stewart D, Hu B, Vinson C, and Alam J. Multiple basic-leucine zipper proteins regulate induction of the mouse heme oxygenase-1 gene by arsenite. Arch Biochem Biophys 405: 265-274, 2002.( @6 R  r4 Y% d. F) n
: W- V9 b% ?* J0 \: Y7 F
! s" x4 @/ L& G# ~- c3 X
+ s6 }) c- g2 x9 n* ~- ?3 i
Hancock WW, Buelow R, Sayegh MH, and Turka LA. Antibody-induced transplant arteriosclerosis is prevented by graft expression of anti-oxidant and anti-apoptotic genes. Nat Med 4: 1392-1396, 1998.7 o$ B0 b% z. ~- D# \2 U
8 J1 B# [2 E& F- Q& x9 s) @
9 u" W& n% K! B  @+ M# G1 P; P9 _
& T: W: C3 s. Q; p& i: F
Hangaishi M, Ishizaka N, Aizawa T, Kurihara Y, Taguchi J, Nagai R, Kimura S, and Ohno M. Induction of heme oxygenase-1 can act protectively against cardiac ischemia/reperfusion in vivo. Biochem Biophys Res Commun 279: 582-588, 2000.) ^0 `- V# f, K3 h

, O3 \, d: t9 _: O# S6 m. i# ?- q7 w/ ]0 ^
* w+ J& i# a. g! g8 ?: c* w
Hara E, Takahashi K, Takeda K, Nakayama M, Yoshizawa M, Fujita H, Shirato K, and Shibahara S. Induction of heme oxygenase-1 as a response in sensing the signals evoked by distinct nitric oxide donors. Biochem Pharmacol 58: 227-236, 1999.' e: [! k" p$ B) l. A( r1 b- J/ ?2 j+ f
) C# S/ M6 _8 ~0 K7 u4 Z: y

+ O0 u7 {: q! V4 S' g
1 O3 C: p3 c- [) dHartsfield CL, Alam J, and Choi AM. Differential signaling pathways of HO-1 gene expression in pulmonary and systemic vascular cells. Am J Physiol Lung Cell Mol Physiol 277: L1133-L1141, 1999.4 ?- F0 S9 \7 O: i/ b4 g; d

' Z6 U  d  |0 U
+ ~, X8 I: Z$ f" \$ s* c. L; |7 y7 l$ N0 O4 r
Hartsfield CL, Alam J, Cook JL, and Choi AM. Regulation of heme oxygenase-1 gene expression in vascular smooth muscle cells by nitric oxide. Am J Physiol Lung Cell Mol Physiol 273: L980-L988, 1997.6 X2 d: a$ M- e+ v+ r

* w4 A7 a9 v$ Q: p+ ]4 \. y# `4 `
+ V! W$ A6 W# |+ s& _2 d' _. X# Q) y0 c) Q% h7 y
Hata A, Seoane J, Lagna G, Montalvo E, Hemmati-Brivanlou A, and Massague J. OAZ uses distinct DNA- and protein-binding zinc fingers in separate BMP-Smad and Olf signaling pathways. Cell 100: 229-240, 2000.$ `6 o9 |3 T( S+ }
8 M, J/ d& g; g2 u9 g* w) d

: `6 c8 x3 s, V% v
* p. H" V4 Q! o1 S2 _Haugen EN, Croatt AJ, and Nath KA. Angiotensin II induces renal oxidant stress in vivo and heme oxygenase-1 in vivo and in vitro. Kidney Int 58: 144-152, 2000.
' |6 c+ J/ p  e" ~8 f. w; m) D2 ^. y6 m2 W" O+ U

7 I. w6 ]7 a1 v0 B: L% G) G% F0 ?9 h2 j6 y" Y) s0 K7 d) ]
Hayashi H, Abdollah S, Qiu Y, Cai J, Xu YY, Grinnell BW, Richardson MA, Topper JN, Gimbrone MA Jr, Wrana JL, and Falb D. The MAD-related protein Smad7 associates with the TGF- receptor and functions as an antagonist of TGF- signaling. Cell 89: 1165-1173, 1997.7 t0 R7 q4 R" O8 N

$ k# c" A0 v: e" Q( }% G
3 O7 Z& F" O  i( c; H# A- B. L1 F6 D9 q+ {- S
He CH, Gong P, Hu B, Stewart D, Choi ME, Choi AM, and Alam J. Identification of activating transcription factor 4 (ATF4) as an Nrf2-interacting protein. Implication for heme oxygenase-1 gene regulation. J Biol Chem 276: 20858-20865, 2001.0 y  ^( G0 U5 _$ B9 D9 @& D

, J! H+ L) V; J0 X, c
7 S/ @! M% @$ Y7 p' b' p0 f1 J8 A* x
0 ~1 e" a) f2 ?7 CHill-Kapturczak N, Chang SH, and Agarwal A. Heme oxygenase and the kidney. DNA Cell Biol 21: 307-321, 2002.+ i) d& i3 X1 e9 N! Q$ b1 u5 K
- q" c3 `4 s* q' e/ E
& f$ s( S( A* `! x" W
# M. A3 J, c! P' t/ p
Hill-Kapturczak N, Sikorski E, Voakes C, Garcia J, Nick HS, and Agarwal A. An internal enhancer regulates heme- and cadmium-mediated induction of human heme oxygenase-1. Am J Physiol Renal Physiol 285: F515-F523, 2003.
7 Z- t3 u" e5 r; n0 W
/ n1 d  {8 v2 \+ @& J( N6 [0 K1 y6 [

* K0 ^, [. P  f* a1 }- z- g( F% e1 |Hill-Kapturczak N, Truong L, Thamilselvan V, Visner GA, Nick HS, and Agarwal A. Smad7-dependent regulation of heme oxygenase-1 by transforming growth factor- in human renal epithelial cells. J Biol Chem 275: 40904-40909, 2000.
3 o: r' s" c7 |7 X% z) `
0 {1 w' K- g# b0 K( u4 U. D6 i& ]+ x! S4 Z& e# {

$ A- q' V* a. l, D# Y: ?Hill-Kapturczak N, Voakes C, Garcia J, Visner G, Nick HS, and Agarwal A. A cis-acting region regulates oxidized lipid-mediated induction of the human heme oxygenase-1 gene in endothelial cells. Arterioscler Thromb Vasc Biol 23: 1416-1422, 2003.  a/ Q" {! J5 z/ b
& I' o. W7 ]& l! g

9 O7 {( `, M/ ]" r) c4 r
5 ^0 b" c$ y: PHorikawa S, Ito K, Ikeda S, Shibata T, Ishizuka S, Yano T, Hagiwara K, Ozasa H, and Katsuyama I. Induction of heme oxygenase-1 in toxic renal injury: mercuric chloride-induced acute renal failure in rat. Toxicol Lett 94: 57-64, 1998.
9 ~2 p7 l. e% K5 ]1 g6 \1 K7 z) ?1 r4 M
* c# {- x5 Y% C. a% D- T( ^2 m& R/ i
, F# {2 _6 P* m" V$ n
Immenschuh S, Hinke V, Katz N, and Kietzmann T. Transcriptional induction of heme oxygenase-1 gene expression by okadaic acid in primary rat hepatocyte cultures. Mol Pharmacol 57: 610-618, 2000.
2 _& w$ G* r- C7 k; d) K* n' _; @$ C" _: }( r! d6 B
7 I. z& f8 Z! q- d
, G& \" b0 a2 `& m6 b
Immenschuh S and Ramadori G. Gene regulation of heme oxygenase-1 as a therapeutic target. Biochem Pharmacol 60: 1121-1128, 2000.+ J, W4 U; Q3 ~: g7 }" `( R
" S( a/ o( U0 T# x: w  N
$ R7 k" S) `4 T5 X2 j

( `: p/ R9 @. O/ [" Y2 ~3 @) `( {) yInamdar NM, Ahn YI, and Alam J. The heme-responsive element of the mouse heme oxygenase-1 gene is an extended AP-1 binding site that resembles the recognition sequences for MAF and NF-E2 transcription factors. Biochem Biophys Res Commun 221: 570-576, 1996.. P& t! w$ t/ C* L0 r. c
: H' l9 ?6 m% |2 q& J

" Y) e* E7 V% k6 Z) ^/ f5 |
1 Y0 I* f8 P, g, S# b4 W# cInguaggiato P, Gonzalez-Michaca L, Croatt AJ, Haggard JJ, Alam J, and Nath KA. Cellular overexpression of heme oxygenase-1 up-regulates p21 and confers resistance to apoptosis. Kidney Int 60: 2181-2191, 2001.9 ~6 z4 B4 Q, C$ n; Q% t
3 L3 l+ f9 W+ K9 p6 V! A5 n
+ c/ G# D! S2 X5 Z4 P' O  {

; T; h1 v/ x0 P1 VIshikawa K, Navab M, Leitinger N, Fogelman AM, and Lusis AJ. Induction of heme oxygenase-1 inhibits the monocyte transmigration induced by mildly oxidized LDL. J Clin Invest 100: 1209-1216, 1997.8 I2 }+ j: g! _+ U+ C

: E) w* [: P+ m
- }! @, I# D1 j, `+ G) N! V2 Z; B9 l+ H# R
Ishikawa K, Sugawara D, Wang X, Suzuki K, Itabe H, Maruyama Y, and Lusis AJ. Heme oxygenase-1 inhibits atherosclerotic lesion formation in ldl-receptor knockout mice. Circ Res 88: 506-512, 2001.
+ X  Z7 ~$ x  r6 K( o' K: c; j: i: I8 Z  p6 v; {
# Y" `4 G3 P3 B! L

  U* n7 @. d6 y' W# X# R$ MIshizaka N, Aizawa T, Mori I, Taguchi J, Yazaki Y, Nagai R, and Ohno M. Heme oxygenase-1 is upregulated in the rat heart in response to chronic administration of angiotensin II. Am J Physiol Heart Circ Physiol 279: H672-H678, 2000.3 y- Y" X9 Z, @% b
) B/ ~  e& s0 L& g, q. b" t
: u3 K9 r  S! J/ W8 ^0 o
6 B$ V& m" y* r/ N7 A  e! p7 `* P
Ishizaka N, Aizawa T, Ohno M, Usui Si S, Mori I, Tang SS, Ingelfinger JR, Kimura S, and Nagai R. Regulation and localization of HSP70 and HSP25 in the kidney of rats undergoing long-term administration of angiotensin II. Hypertension 39: 122-128, 2002.
7 U0 ^( Y) m  m
4 t) H9 d- C  \4 V0 `" M' I" D# T( h/ M- g3 C* \) E

/ C) J& ^4 u1 KIshizaka N, Aizawa T, Yamazaki I, Usui S, Mori I, Kurokawa K, Tang SS, Ingelfinger JR, Ohno M, and Nagai R. Abnormal iron deposition in renal cells in the rat with chronic angiotensin II administration. Lab Invest 82: 87-96, 2002.
, q7 y2 L! {& Q3 y' G' O; G4 {2 |" \0 O7 K# j

9 \8 k) n! n+ m# [) ]+ G  ?, U( z7 q% v
Ishizaka N, de Leon H, Laursen JB, Fukui T, Wilcox JN, De Keulenaer G, Griendling KK, and Alexander RW. Angiotensin II-induced hypertension increases heme oxygenase-1 expression in rat aorta. Circulation 96: 1923-1929, 1997.
6 W9 l) e4 {1 ]0 H5 {+ H9 z) k" ^3 O9 |6 |

+ I3 s. E0 f7 y
0 j2 b8 _0 q9 X2 O5 J" ~( M6 OItoh K, Chiba T, Takahashi S, Ishii T, Igarashi K, Katoh Y, Oyake T, Hayashi N, Satoh K, Hatayama I, Yamamoto M, and Nabeshima Y. An Nrf2/small Maf heterodimer mediates the induction of phase II detoxifying enzyme genes through antioxidant response elements. Biochem Biophys Res Commun 236: 313-322, 1997.
0 V3 y; Y- @" {7 N: \) g8 ^' ?& G5 x( C9 \, w  t1 g1 T# ^
  k- z$ M: c: e

2 o0 q  a- D* D! b( p% I% ?0 WKacimi R, Chentoufi J, Honbo N, Long CS, and Karliner JS. Hypoxia differentially regulates stress proteins in cultured cardiomyocytes: role of the p38 stress-activated kinase signaling cascade, and relation to cytoprotection. Cardiovasc Res 46: 139-150, 2000.
- s2 h  B5 X8 @* b4 u* i1 x5 h5 P/ g$ Q" j- e. X  v
6 V1 X- i1 B! `; ^. J0 O

0 \* C7 s* _8 U. s# i% dKaide JI, Zhang F, Wei Y, Jiang H, Yu C, Wang WH, Balazy M, Abraham NG, and Nasjletti A. Carbon monoxide of vascular origin attenuates the sensitivity of renal arterial vessels to vasoconstrictors. J Clin Invest 107: 1163-1171, 2001.
( h. [  G1 c$ A* n: ?# V; w5 H5 U: X$ a% M- I( @; F

8 K' p% J: O$ m+ Q# [/ Y: q5 W) n0 W$ j0 g
Kaneda H, Ohno M, Taguchi J, Togo M, Hashimoto H, Ogasawara K, Aizawa T, Ishizaka N, and Nagai R. Heme oxygenase-1 gene promoter polymorphism is associated with coronary artery disease in Japanese patients with coronary risk factors. Arterioscler Thromb Vasc Biol 22: 1680-1685, 2002.
( A7 I9 D$ Z6 X( L, q6 h; o( o, k$ B$ q

% O8 \0 `' V5 N; X) P1 D
" N: M" ?5 K2 H9 N# d- qKanwar YS. Heme oxygenase-1 in renal injury: conclusions of studies in humans and animal models. Kidney Int 59: 378-379, 2001.  R; N: D& n3 ^* e% J9 U
& Q# G- `5 K& q  p7 Y

0 n2 W! W  h& ]! O$ _- X5 I6 r2 x- [2 x0 J
Kataoka K, Handa H, and Nishizawa M. Induction of cellular antioxidative stress genes through heterodimeric transcription factor Nrf2/small Maf by antirheumatic gold(I) compounds. J Biol Chem 276: 34074-34081, 2001.* d2 o1 S( C3 L4 ?+ G! ]8 v

  X& W( a5 l8 B4 @0 t
5 [- F- z6 g2 ^2 {" v- e$ h9 D$ N' f4 G" w% n2 j9 l# H2 G4 ]3 X
Katoh Y, Itoh K, Yoshida E, Miyagishi M, Fukamizu A, and Yamamoto M. Two domains of Nrf2 cooperatively bind CBP, a CREB binding protein, and synergistically activate transcription. Genes Cells 6: 857-868, 2001.
  v& z: `/ j  N( c3 ~" K& N
6 E$ |  r, e% y* x/ a  b' o' P& p: G& j4 R) g0 m' a( g

! L6 y" \: N" G7 o$ X; U* V; eKatori M, Busuttil RW, and Kupiec-Weglinski JW. Heme oxygenase-1 system in organ transplantation. Transplantation 74: 905-912, 2002.
; B1 Y0 p' U) ?  {+ B! O4 Y1 z; R8 z1 ~' U: V
4 z4 @* {* T1 }$ W8 L
$ G# z" I3 `/ U6 u  c% u
Kays SE, Nowak G, and Schnellmann RG. Transforming growth factor- 1 inhibits regeneration of renal proximal tubular cells after oxidant exposure. J Biochem Toxicol 11: 79-84, 1996. <a href="/cgi/external_ref?access_num=10.1002/(SICI)1522-7146(1996)11:23 ]+ U; U" l1 J/ c0 [  z* o, w3 |

( W" K' P2 g3 ~/ g3 b, |: G- M# k, o/ I0 [' Y; N
* F$ B; [5 ^# l% A# A2 S: J
Ke B, Shen XD, Zhai Y, Gao F, Busuttil RW, Volk HD, and Kupiec-Weglinski JW. Heme oxygenase 1 mediates the immunomodulatory and antiapoptotic effects of interleukin 13 gene therapy in vivo and in vitro. Hum Gene Ther 13: 1845-1857, 2002.
4 G5 g' ?( x$ v# K% g! j2 a
! X; o- R; h4 c. Y/ j
( \$ o) E- [. U9 r: I  A  D6 q- @; V+ K$ y$ O" }$ C' p+ z
Keyse SM and Tyrrell RM. Heme oxygenase is the major 32-kDa stress protein induced in human skin fibroblasts by UVA radiation, hydrogen peroxide, and sodium arsenite. Proc Natl Acad Sci USA 86: 99-103, 1989.
2 v1 U! t; k9 @3 u
7 G  C7 W4 J4 i8 H! \2 z8 B2 S6 Q
6 B$ X6 n6 {6 k5 X8 _5 S8 N0 d" P7 f, j$ w9 ^# }) a+ j
Kietzmann T, Samoylenko A, and Immenschuh S. Transcriptional regulation of heme oxygenase-1 gene expression by MAP kinases of the JNK and p38 pathways in primary cultures of rat hepatocytes. J Biol Chem 278: 17927-17936, 2003.
! n1 c+ H( }, b+ Q& `/ N
5 h" k0 s* a8 F" I5 p/ f, R) [, E) \. A

+ ~% R' Q# q. ]Kitamuro T, Takahashi K, Ogawa K, Udono-Fujimori R, Takeda K, Furuyama K, Nakayama M, Sun J, Fujita H, Hida W, Hattori T, Shirato K, Igarashi K, and Shibahara S. Bach1 functions as a hypoxiainducible repressor for the heme oxygenase-1 gene in human cells. J Biol Chem 278: 9125-9133, 2003.
( ^5 g1 i  Y. [# G9 E) u
0 O# W% [. H1 ?4 ^) ^" U6 q) o  _0 ]5 P

8 u: t  x6 b# ^' c) D; E5 CKozma F, Johnson RA, and Nasjletti A. Role of carbon monoxide in heme-induced vasodilation. Eur J Pharmacol 323: R1-R2, 1997.. T! i: X7 W! O- d: a8 I0 g5 Y

! u8 X9 q% ]  e: V1 i' X+ B$ \
. l* X# [& b0 r/ \; R* B/ E# s8 P4 y! F& h
Kurata S, Matsumoto M, Tsuji Y, and Nakajima H. Lipopolysaccharide activates transcription of the heme oxygenase gene in mouse M1 cells through oxidative activation of nuclear factor B. Eur J Biochem 239: 566-571, 1996.
4 j( h9 W: E' r( q  y' h8 f- i( E. F9 F- J! t+ `" e
3 G$ B6 n, H, M& K4 B5 I3 Q

- W6 o. s; i( x9 NKutty RK, Kutty G, Rodriguez IR, Chader GJ, and Wiggert B. Chromosomal localization of the human heme oxygenase genes: heme oxygenase-1 (HMOX1) maps to chromosome 22q12 and heme oxygenase-2 (HMOX2) maps to chromosome 16p13.3. Genomics 20: 513-516, 1994.
' B5 O3 p/ @% _) Q: ]+ [# Q' ?7 V* R0 g

2 M; N% h& T/ M' t3 j$ d& U, F- l! o
" q! m0 E: p+ TKutty RK, Nagineni CN, Kutty G, Hooks JJ, Chader GJ, and Wiggert B. Increased expression of heme oxygenase-1 in human retinal pigment epithelial cells by transforming growth factor-beta. J Cell Physiol 159: 371-378, 1994.; q$ v- T8 A# B

! s, I  x2 R- g; C, \9 i4 Z5 z& v0 g
" h9 s0 n! W$ C% F
Lavrovsky Y, Schwartzman ML, Levere RD, Kappas A, and Abraham NG. Identification of binding sites for transcription factors NF- B and AP-2 in the promoter region of the human heme oxygenase 1 gene. Proc Natl Acad Sci USA 91: 5987-5991, 1994.3 x" e* ]+ J3 D; [( W! ?

- m# c& U+ F) `$ I* }5 a3 d* a& k# K3 Y4 }6 P
9 P/ u) n$ M+ P# d" @
Lee PJ, Alam J, Sylvester SL, Inamdar N, Otterbein L, and Choi AM. Regulation of heme oxygenase-1 expression in vivo and in vitro in hyperoxic lung injury. Am J Respir Cell Mol Biol 14: 556-568, 1996.
# G; O, }* Y0 a. L$ r' f$ L1 E
5 l( _/ ^6 ]2 ~7 a8 `4 u4 Y
8 S5 t4 Q3 t+ P+ n" U# ?! C0 L
' x% q* w% Z6 M1 Z8 ?4 }Lee PJ, Jiang BH, Chin BY, Iyer NV, Alam J, Semenza GL, and Choi AM. Hypoxia-inducible factor-1 mediates transcriptional activation of the heme oxygenase-1 gene in response to hypoxia. J Biol Chem 272: 5375-5381, 1997.  J5 t/ ]) ^0 i* v
' h+ [  G0 `9 y7 c: X3 ]" c& g# F) m
7 E( M5 C2 h3 j9 Y' H5 j9 n
: \5 i/ D" A, I- W
Lee TS and Chau LY. Heme oxygenase-1 mediates the anti-inflammatory effect of interleukin-10 in mice. Nat Med 8: 240-246, 2002.
4 J0 P1 v, o( w7 o9 p) X. l
. e# K. L$ \6 H3 i( {% E) U% A2 u6 x' {
, k* x" n) z$ g1 r; k# ?. k3 ?
, }. ~- G. D/ z; X2 @Li N, Venkatesan MI, Miguel A, Kaplan R, Gujuluva C, Alam J, and Nel A. Induction of heme oxygenase-1 expression in macrophages by diesel exhaust particle chemicals and quinones via the antioxidant-responsive element. J Immunol 165: 3393-3401, 2000.' D2 M; G; d5 \' C
7 R% U# C5 @1 T! w+ l& }  a
6 e$ V3 p% o7 u
/ e4 M& u% R! G# K4 j. o7 t" W
Liang M, Croatt AJ, and Nath KA. Mechanisms underlying induction of heme oxygenase-1 by nitric oxide in renal tubular epithelial cells. Am J Physiol Renal Physiol 279: F728-F735, 2000.
7 D" y' l* J' v$ q5 R/ f* k; s3 O
. f6 F4 q2 x9 G6 Z5 u

+ G4 F, _# X. y+ x9 K- i* I" FLing H, Edelstein C, Gengaro P, Meng X, Lucia S, Knotek M, Wangsiripaisan A, Shi Y, and Schrier R. Attenuation of renal ischemia-reperfusion injury in inducible nitric oxide synthase knockout mice. Am J Physiol Renal Physiol 277: F383-F390, 1999.
) Y3 x. Q+ `' N: t( J& y; N4 B0 H# }  z9 o; n, F7 f

9 O" f# Y1 l" e$ `+ M4 n; ~7 @- x+ ]5 n) O) t. P
Ling H, Gengaro PE, Edelstein CL, Martin PY, Wangsiripaisan A, Nemenoff R, and Schrier RW. Effect of hypoxia on proximal tubules isolated from nitric oxide synthase knockout mice. Kidney Int 53: 1642-1646, 1998.
0 E& a/ \8 d0 \! o
1 Q7 S1 k% J  n6 h& X1 l
: `' h. q: `5 o/ ?  B9 v! ~: ~5 D
4 M, |& U8 s( M/ N* @Llesuy SF and Tomaro ML. Heme oxygenase and oxidative stress. Evidence of involvement of bilirubin as physiological protector against oxidative damage. Biochim Biophys Acta 1223: 9-14, 1994.
. A# G- l+ K! \/ k' Q! b  |) ~3 [" e. x$ Y+ p  W

. W6 g1 H( W; V( {; T4 w1 A0 U  A, {4 R! a, F% O/ W
Lordnejad MR, Schliess F, Wettstein M, and Haussinger D. Modulation of the heme oxygenase HO-1 expression by hyperosmolarity and betaine in primary rat hepatocytes. Arch Biochem Biophys 388: 285-292, 2001.
, R# Q# Q3 Z. ~9 f
, T! ?4 n# l1 R: D4 J
. Q' a* @* ]% Z7 I4 }! Q
2 W; P) i* V' l3 @; A: _Lu TH, Lambrecht RW, Pepe J, Shan Y, Kim T, and Bonkovsky HL. Molecular cloning, characterization, and expression of the chicken heme oxygenase-1 gene in transfected primary cultures of chick embryo liver cells. Gene 207: 177-186, 1998./ q; H2 t  ~: X( d
) }1 |. h5 k, ]% P7 X

) }3 O! \/ n/ ]) ~2 L
4 q6 \( w0 b. |. d7 s! C) kMaeshima H, Sato M, Ishikawa K, Katagata Y, and Yoshida T. Participation of altered upstream stimulatory factor in the induction of rat heme oxygenase-1 by cadmium. Nucleic Acids Res 24: 2959-2965, 1996.
$ I' N# b5 }# ^- B/ ?3 h* g' [5 R" d
( m: r& c. C5 V2 |/ S0 ]0 J) {/ ~+ B6 }2 b/ I2 w& \2 S  N# S

7 i1 O9 Q# I: k% q. _2 c# zMagee CC, Azuma H, Knoflach A, Denton MD, Chandraker A, Iyer S, Buelow R, and Sayegh M. In vitro and in vivo immunomodulatory effects of RDP1258, a novel synthetic peptide. J Am Soc Nephrol 10: 1997-2005, 1999.
/ y5 j; W8 D4 b+ J( K2 }. m
! T6 m9 u% B6 P1 ?: R0 Z8 o/ Q
4 J- ~. B4 Z3 n% t5 e" X6 J6 d- j. G3 H( D) z* u
Maines MD. The heme oxygenase system: a regulator of second messenger gases. Annu Rev Pharmacol Toxicol 37: 517-554, 1997.5 n6 g" G; g+ c% A7 @6 u
3 f$ H) m/ m7 t
+ g/ d& C! h$ H. X: v9 k+ `

0 y1 {3 C. Z% W) N/ EMaines MD, Mayer RD, Ewing JF, and McCoubrey WK Jr. Induction of kidney heme oxygenase-1 (HSP32) mRNA and protein by ischemia/reperfusion: possible role of heme as both promotor of tissue damage and regulator of HSP32. J Pharmacol Exp Ther 264: 457-462, 1993.
0 ~" ~' y4 h2 {/ Z& I4 T
( {! Y4 p% u" `; d- k2 L# B( @4 }; _; r  C0 Y7 v
9 c! [1 t' W! m( }
Marini MG, Chan K, Casula L, Kan YW, Cao A, and Moi P. hMAF, a small human transcription factor that heterodimerizes specifically with Nrf1 and Nrf2. J Biol Chem 272: 16490-16497, 1997.6 Q) {, F! h, u6 Y  e8 N( O7 E6 W+ |
5 O0 P0 `- r  w6 h% u4 L

0 `- G8 ]$ g7 @) X/ A  q1 F# T$ P
Marks GS, Brien JF, Nakatsu K, and McLaughlin BE. Does carbon monoxide have a physiological function? Trends Pharmacol Sci 12: 185-188, 1991.% c* l' t2 g$ X3 X

, z8 n; M! c7 ?( ]* ]/ {' C: l3 j& j6 v! p9 n6 U
4 C" i7 t! ?' e4 ]
Marquis JC and Demple B. Complex genetic response of human cells to sublethal levels of pure nitric oxide. Cancer Res 58: 3435-3440, 1998.
& M  x; s) {; `7 v; p. f
' N* Q0 D) E" m- i9 p" u- j% T5 f, _- F8 j# J6 n5 i( F( R
. ~$ W& ~" U7 F3 R  y: `' Z# W
Massague J. How cells read TGF- signals. Nat Rev Mol Cell Biol 1: 169-178, 2000.  H* r  B/ D# B1 Q3 D  P6 z9 O

0 `7 a; R1 R) ^3 j2 D
( u& R: w5 k+ ^2 l
& B# J  P8 L3 @8 q5 jMasuya Y, Hioki K, Tokunaga R, and Taketani S. Involvement of the tyrosine phosphorylation pathway in induction of human heme oxygenase-1 by hemin, sodium arsenite, and cadmium chloride. J Biochem (Tokyo) 124: 628-633, 1998.* ?2 c5 o4 E4 U+ v! |& O
# B/ `; b7 e9 k3 ]- A

- ^% ^6 b4 ~0 R
0 M, M( \3 t7 c+ AMazza F, Goodman A, Lombardo G, Vanella A, and Abraham NG. Heme oxygenase-1 gene expression attenuates angiotensin II-mediated DNA damage in endothelial cells. Exp Biol Med (Maywood) 228: 576-583, 2003.
1 V; c3 n7 ]6 r  M9 \: ~4 [; r. F+ z; i2 Y

; P/ ^4 P) [+ M% w  g+ j( x( _. E
6 Z2 x3 L+ A8 h, ]% F9 b' f' xMcCoubrey WK Jr, Huang TJ, and Maines MD. Isolation and characterization of a cDNA from the rat brain that encodes hemoprotein heme oxygenase-3. Eur J Biochem 247: 725-732, 1997.
/ D! m6 w% l7 Q; m% q
6 V4 l% e+ W& i; Q" C; v  L0 |: y* [( m+ T0 ?. z+ E, x
7 U5 ]2 Z. }# k- v0 w# T$ T6 @
Mellott JK, Nick HS, Waters MF, Billiar TR, Geller DA, and Chesrown SE. Cytokine-induced changes in chromatin structure and in vivo footprints in the inducible NOS promoter. Am J Physiol Lung Cell Mol Physiol 280: L390-L399, 2001.1 V5 k' X8 ]! N/ R$ f: v- q* ^
( G) Y6 e( o$ X- U  `

* B% }. ~$ Y8 t& @3 {* m2 K, r; V# @" Y7 z( R: c' |
Mitani K, Fujita H, Kappas A, and Sassa S. Heme oxygenase is a positive acute-phase reactant in human Hep3B hepatoma cells. Blood 79: 1255-1259, 1992.  L9 G8 S& a! ]

' S3 i# j  P# F9 h  C
* E: I7 N7 F$ Q( A" g
& M' b2 s' ]1 e0 p& qMoi P, Chan K, Asunis I, Cao A, and Kan YW. Isolation of NF-E2-related factor 2 (Nrf2), a NF-E2-like basic leucine zipper transcriptional activator that binds to the tandem NF-E2/AP1 repeat of the -globin locus control region. Proc Natl Acad Sci USA 91: 9926-9930, 1994.; W# ]! l- X5 Q+ t; ~) [
) g0 ?. N- B% H/ P& D) n7 j8 X: T
# ~% w0 D. X2 L' p/ a7 i, Y

  O1 W: @% L0 b5 i* j  |Montellano PR. The mechanism of heme oxygenase. Curr Opin Chem Biol 4: 221-227, 2000.
: k" m* @' ~2 t" r1 d8 \
) ?7 e( Z) W4 N0 B( {+ }/ A8 e- I* x! n# L- X" s% m0 ~3 `
) G# S4 u! Y4 c% @
Mosley K, Wembridge DE, Cattell V, and Cook HT. Heme oxygenase is induced in nephrotoxic nephritis and hemin, a stimulator of heme oxygenase synthesis, ameliorates disease. Kidney Int 53: 672-678, 1998.5 o/ h. G: U6 n2 R
: }& k" A" U. c1 k8 H/ V
* S( L+ g; G6 n) |8 {; `

/ K" j, H: M" j- eMotohashi H, O'Connor T, Katsuoka F, Engel JD, and Yamamoto M. Integration and diversity of the regulatory network composed of Maf and CNC families of transcription factors. Gene 294: 1-12, 2002." p( n# J/ o4 L3 Y: h: I

/ a. C! |! i  `7 q: y
5 U. h1 s/ e! K( O" Y, J- R# S4 X+ W" g# k& Q1 w. [: P% r; d8 X
Motterlini R, Foresti R, Bassi R, Calabrese V, Clark JE, and Green CJ. Endothelial heme oxygenase-1 induction by hypoxia. Modulation by inducible nitric-oxide synthase and S -nitrosothiols. J Biol Chem 275: 13613-13620, 2000.$ G) d* Y0 V$ {5 r1 J9 k4 P( p
! z: G' ^3 [, u0 ^( l/ S/ d2 u

) w8 N' T# q/ r% |  d
% S- q# x, z1 p! }+ XMotterlini R, Foresti R, Intaglietta M, and Winslow RM. NO-mediated activation of heme oxygenase: endogenous cytoprotection against oxidative stress to endothelium. Am J Physiol Heart Circ Physiol 270: H107-H114, 1996.
8 u5 \1 S' s, M3 F
  s# F/ N- y/ i( H2 W% b2 |
/ u2 `# @/ {" f3 t& b
/ E) O; H$ r" s' ^- C1 yMuller RM, Taguchi H, and Shibahara S. Nucleotide sequence and organization of the rat heme oxygenase gene. J Biol Chem 262: 6795-6802, 1987./ }# }( Q& c: n( k% d
- q0 w8 c" l0 n
" m; y  T; ~% C7 f0 o3 m6 w+ k2 y

9 O  y6 T  g8 f* C" |& mMuraosa Y, Takahashi K, Yoshizawa M, and Shibahara S. cDNA cloning of a novel protein containing two zinc-finger domains that may function as a transcription factor for the human heme-oxygenase-1 gene. Eur J Biochem 235: 471-479, 1996.1 u5 R; I2 Y) @4 u6 t6 H
: B/ S) z7 C# j

: M; m1 n4 B1 Z9 Z5 \6 r' d% A" W& F) D: }# `
Nagarajan RP, Zhang J, Li W, and Chen Y. Regulation of Smad7 promoter by direct association with Smad3 and Smad4. J Biol Chem 274: 33412-33418, 1999.
+ L1 E4 `7 I0 O7 }: }
' `& m9 A) B, u
5 f; ^6 ^- v' Y! @0 q% u7 |2 S: N  J& Y; R) z' e
Nakao A, Fujii M, Matsumura R, Kumano K, Saito Y, Miyazono K, and Iwamoto I. Transient gene transfer and expression of Smad7 prevents bleomycin-induced lung fibrosis in mice. J Clin Invest 104: 5-11, 1999.
, u/ t8 c& G/ t
7 |# @( M0 |* ^  O: r) s( x! O* `* `8 N$ Q  _( V

$ z" D7 x! W2 b  k, K# ANath KA. Heme oxygenase-1: a redoubtable response that limits reperfusion injury in the transplanted adipose liver. J Clin Invest 104: 1485-1486, 1999.
" W) A# g) S/ M+ t& o2 M3 M
- Y! v* {0 |! t# X! A, H4 [, S& P' D$ x, }8 O
/ l9 z" P. R' Y! {# L+ h2 I
Nath KA, Balla G, Vercellotti GM, Balla J, Jacob HS, Levitt MD, and Rosenberg ME. Induction of heme oxygenase is a rapid, protective response in rhabdomyolysis in the rat. J Clin Invest 90: 267-270, 1992.
1 ~! b# M+ \: k2 ~; a/ H% g5 z  X! R9 ?$ n1 u0 m& t+ {# g- a

7 G6 y1 D+ w5 [8 M# C/ z0 H( {& ?- r+ }. q4 T" z9 A& s
Nath KA, Croatt AJ, Likely S, Behrens TW, and Warden D. Renal oxidant injury and oxidant response induced by mercury. Kidney Int 50: 1032-1043, 1996.! `. M8 ~# V" l" |
  M" i/ e+ Q5 P2 @4 m* D: h
. R, i' E/ n: M$ T3 y  k! ?5 }4 \2 i

6 l% V' H! F3 G/ |) r) m+ O" f9 HNath KA, Grande JP, Croatt AJ, Likely S, Hebbel RP, and Enright H. Intracellular targets in heme protein-induced renal injury. Kidney Int 53: 100-111, 1998.
1 [2 Q8 ]8 @; P1 P7 z: w7 X4 U8 J% e2 Z# p8 Z5 L2 S6 w+ N. q
" W1 e5 ?6 q( X: S" i) C
6 }! A/ v# h& b" d) V* j
Nath KA, Grande JP, Haggard JJ, Croatt AJ, Katusic ZS, Solovey A, and Hebbel RP. Oxidative stress and induction of heme oxygenase-1 in the kidney in sickle cell disease. Am J Pathol 158: 893-903, 2001.! `+ _& N# Q7 t

" t3 z$ ?1 s" X( g1 Y. X/ _
8 m2 |+ x8 W. M& l+ T2 ?8 W# A( w* r: J- j! {0 @" W
Nath KA, Haggard JJ, Croatt AJ, Grande JP, Poss KD, and Alam J. The indispensability of heme oxygenase-1 in protecting against acute heme protein-induced toxicity in vivo. Am J Pathol 156: 1527-1535, 2000.0 p, i3 G* G% }  D: G7 ]) n& O

2 g9 j" L) L' m* A+ f0 J6 a* _/ e5 F+ {0 M' e8 A* k' Z) f0 Y; s# N

$ B. ~' Q) ~0 A" \. x+ Q& ^6 D- lNath KA and Norby SM. Reactive oxygen species and acute renal failure. Am J Med 109: 665-678, 2000.6 N0 ?6 D" @  j! E

8 u' z$ I" Z9 m
  w# K) V  l$ F) |& a0 I+ M1 B
2 u5 u6 R' e" K% H7 rNath KA, Vercellotti GM, Grande JP, Miyoshi H, Paya CV, Manivel JC, Haggard JJ, Croatt AJ, Payne WD, and Alam J. Heme protein-induced chronic renal inflammation: suppressive effect of induced heme oxygenase-1. Kidney Int 59: 106-117, 2001.
3 ^5 F0 S# }3 S5 y# ~; O1 v' ?$ [* B  Y5 k. z+ y" [
0 T  b7 S& C- \2 g* P6 ^
5 k7 U1 k4 N5 H5 l' f
Naylor LH and Clark EM. d(TG)n.d(CA)n sequences upstream of the rat prolactin gene form Z-DNA and inhibit gene transcription. Nucleic Acids Res 18: 1595-1601, 1990.
& l# T% Z0 ^* K; |5 {6 Q  H8 w( O$ `  y, U. K( T

4 B; X. q7 i! k& _: n4 _! I5 h- B' O% v& }2 l6 O3 \* i
Nguyen T, Sherratt PJ, and Pickett CB. Regulatory mechanisms controlling gene expression mediated by the antioxidant response element. Annu Rev Pharmacol Toxicol 43: 233-260, 2003.
2 \8 r. u$ v, q7 [( o6 f8 F+ Q7 u0 S) {
. M* ?# F! k" O" _& p

; \6 s1 ~8 C) pNigam S and Lieberthal W. Acute renal failure. III. The role of growth factors in the process of renal regeneration and repair. Am J Physiol Renal Physiol 279: F3-F11, 2000.
: v- L! \5 W. F! n/ `' T) B7 ?8 h
6 s8 I3 C* p' {- q& ^2 E/ J  A" E8 @8 G/ q$ T" L
8 g! y% x8 v# P" \1 H
Nimura T, Weinstein PR, Massa SM, Panter S, and Sharp FR. Heme oxygenase-1 (HO-1) protein induction in rat brain following focal ischemia. Brain Res Mol Brain Res 37: 201-208, 1996.
& I! U9 k4 K- Z2 ]( Q) V7 e- d  q' X$ _9 W6 P/ Y% u3 ^

8 q5 v" ]! p, @1 j* n" [! s+ b
" A6 G) Y  u4 x  p- \Ning W, Song R, Li C, Park E, Mohsenin A, Choi AM, and Choi ME. TGF- 1 stimulates HO-1 via the p38 mitogen-activated protein kinase in A549 pulmonary epithelial cells. Am J Physiol Lung Cell Mol Physiol 283: L1094-L1102, 2002.
7 N! _2 y/ l9 s9 d2 B2 M5 A9 m: q. o, u$ e, A8 t8 `1 h; X
  {2 N' n) s3 n
4 j4 J/ ^+ R% ^) K# E9 v& m
Nishioka Y and Leder P. Organization and complete sequence of identical embryonic and plasmacytoma kappa V-region genes. J Biol Chem 255: 3691-3694, 1980.
4 h" U6 S+ e& }" {* i# w
9 U" F' G% r5 b; t/ N
6 K; s3 Y4 c0 U  a& [- _
8 c+ Z. m) r* D' ]) FNoble NA and Border WA. Angiotensin II in renal fibrosis: should TGF- rather than blood pressure be the therapeutic target? Semin Nephrol 17: 455-466, 1997., r& I' W! D  T6 Q
! E' t1 \: [: \9 K. ^

- e/ s( S) @, O9 a8 L
3 ]# b& `' D' |4 CNoiri E, Peresleni T, Miller F, and Goligorsky MS. In vivo targeting of inducible NO synthase with oligodeoxynucleotides protects rat kidney against ischemia. J Clin Invest 97: 2377-2383, 1996.
: r" W5 D4 i$ ~3 I. i* Y: K8 g- i2 E$ B

' T2 Z8 A6 z5 O. p, S: B; E4 e+ q; r3 ^( p- c" o, K
Noris M. Nitric oxide as a mediator of hemodynamic disturbances in acute renal failure associated with sepsis. In: The Kidney in Sepsis, Critical Care in Nephrology, edited by Ronco C and Bellomo R. Dordrecht, The Netherlands: Kluwer Academic, 1998, p. 575-589.0 c6 m" w/ ~2 _) o
* b7 g- H; k" V$ z) j- N' d
" e* x/ ]; o. B( G; O
, j. s* @; W# e! [& `  m# S) o
Ogawa K, Sun J, Taketani S, Nakajima O, Nishitani C, Sassa S, Hayashi N, Yamamoto M, Shibahara S, Fujita H, and Igarashi K. Heme mediates derepression of Maf recognition element through direct binding to transcription repressor Bach1. EMBO J 20: 2835-2843, 2001.
/ L( q* i# U( U8 @! D  S' x9 d
. S4 Z! p: x" t* Z) M0 N$ H$ P4 S; v7 W( _" ?0 L! C  W( @- P! a# \
, O  c( e& |0 n9 b6 [/ S
Oguro T, Takahashi Y, Ashino T, Takaki A, Shioda S, Horai R, Asano M, Sekikawa K, Iwakura Y, and Yoshida T. Involvement of tumor necrosis factor alpha, rather than interleukin-1 / or nitric oxides in the heme oxygenase-1 gene expression by lipopolysaccharide in the mouse liver. FEBS Lett 516: 63-66, 2002.
9 k( o0 t8 p+ Q" g- d" E4 e: B( L3 c' k2 z. b1 I

4 Y/ Y  Z! K  T" l" g* ~; A
( S& ]3 ?5 B" L( X! F& U1 TOhira M, Seki N, Nagase T, Ishikawa K, Nomura N, and Ohara O. Characterization of a human homolog (BACH1) of the mouse Bach1 gene encoding a BTB-basic leucine zipper transcription factor and its mapping to chromosome 21q22.1. Genomics 47: 300-306, 1998.
+ M' L2 L& E" ~" L' }
* U# |0 t( j7 a
& Q9 t# e" g% N. H3 N0 B2 `# u
2 A, t9 q; n" r; h9 h1 m2 uOhta K, Yachie A, Fujimoto K, Kaneda H, Wada T, Toma T, Seno A, Kasahara Y, Yokoyama H, Seki H, and Koizumi S. Tubular injury as a cardinal pathologic feature in human heme oxygenase-1 deficiency. Am J Kidney Dis 35: 863-870, 2000.
' T2 `- \$ _. E2 y: n
* e3 g: {- S# K2 l% Z& c1 `/ J5 `" Y% s$ c- {

( V" H& i( o- s  `Oken DE, Mende CW, Taraba I, and Flamenbaum W. Resistance to acute renal failure afforded by prior renal failure: examination of the role of renal renin content. Nephron 15: 131-142, 1975.
: H  B6 c. ^9 i; l: j$ e/ y; e) B/ j; G

% z7 i5 A1 Q* f6 M/ T0 U8 K& M% ?+ @/ \8 ~
Okinaga S, Takahashi K, Takeda K, Yoshizawa M, Fujita H, Sasaki H, and Shibahara S. Regulation of human heme oxygenase-1 gene expression under thermal stress. Blood 87: 5074-5084, 1996.% ^+ {; z4 N2 R, U6 G

! f1 x1 T$ x7 `, h5 y, j1 t$ y3 Y" `6 }

. E- Y( }. G, {, \0 EOshiro S, Takeuchi H, Matsumoto M, and Kurata S. Transcriptional activation of heme oxygenase-1 gene in mouse spleen, liver and kidney cells after treatment with lipopolysaccharide or hemoglobin. Cell Biol Int 23: 465-474, 1999.
- V" _5 |/ e  }' I
8 K1 V9 P" Y8 [" q# I
, Q+ g! g, o* O3 c. J
# I: N  ?. z  W- D9 D6 oOtterbein LE, Bach FH, Alam J, Soares M, Tao Lu H, Wysk M, Davis RJ, Flavell RA, and Choi AM. Carbon monoxide has anti-inflammatory effects involving the mitogen-activated protein kinase pathway. Nat Med 6: 422-428, 2000.
/ e2 C- F# \, o. H
% O9 f$ V$ D3 ?' [- {6 k% u
4 O; C- _! q6 \9 h$ u& g, c# I, h' a/ V
Otterbein LE and Choi AM. Heme oxygenase: colors of defense against cellular stress. Am J Physiol Lung Cell Mol Physiol 279: L1029-L1037, 2000.! {$ F# s- J  N0 F+ x1 q+ q7 Y! ^$ Y! y
3 Y, v/ |8 z! w) ^4 Y
' _0 ~1 Z! [. f( N
9 H. v3 u, S% }0 [9 E
Otterbein LE, Mantell LL, and Choi AM. Carbon monoxide provides protection against hyperoxic lung injury. Am J Physiol Lung Cell Mol Physiol 276: L688-L694, 1999.
* B" X4 S8 T, [/ o% o7 z* y: _- k0 u
. g" }5 y" d  s0 N( M- t2 k" M5 a
- M0 K  k0 v3 r5 ]) i, v
Otterbein LE, Soares MP, Yamashita K, and Bach FH. Heme oxygenase-1: unleashing the protective properties of heme. Trends Immunol 24: 449-455, 2003.
4 Z$ a# F: C0 j" [
) L9 J# p/ a9 _5 b
0 _- W) C; J" p9 W+ d  R  F; X! f! V# Z: a9 w
Oyake T, Itoh K, Motohashi H, Hayashi N, Hoshino H, Nishizawa M, Yamamoto M, and Igarashi K. Bach proteins belong to a novel family of BTB-basic leucine zipper transcription factors that interact with MafK and regulate transcription through the NF-E2 site. Mol Cell Biol 16: 6083-6095, 1996.
: {7 v5 Z( `/ x9 t, |( A( P, m1 s( A- k1 P$ r# K; x
' p% D4 z1 \6 I
' D. }/ O% k" a' `) e
Panahian N, Yoshiura M, and Maines MD. Overexpression of heme oxygenase-1 is neuroprotective in a model of permanent middle cerebral artery occlusion in transgenic mice. J Neurochem 72: 1187-1203, 1999.( S, U9 _  c- c* k& w4 O/ p; I
4 E6 D) c9 }4 ~# k  x8 Y

% Y; M9 D+ }' {3 |% u7 q, O( N. ?/ k4 W. g  r
Pellacani A, Wiesel P, Sharma A, Foster LC, Huggins GS, Yet SF, and Perrella MA. Induction of heme oxygenase-1 during endotoxemia is downregulated by transforming growth factor- 1. Circ Res 83: 396-403, 1998.  p0 W* B0 J9 H. ^
4 V+ u0 e; z+ I. c$ A

0 k' S* q: M% q7 s/ I% T* J
+ Z, w1 s+ ^) g3 k- i5 w8 a7 v6 WPimstone NR, Engel P, Tenhunen R, Seitz PT, Marver HS, and Schmid R. Inducible heme oxygenase in the kidney: a model for the homeostatic control of hemoglobin catabolism. J Clin Invest 50: 2042-2050, 1971.
9 o+ j5 q6 b6 w+ u2 }, m0 r% z) {. A- g) E* ?2 {6 Y; `: N
9 `5 G8 W( t) K, }" d$ H% y' C
& S) a) k- n* n. [4 Q
Platt JL and Nath KA. Heme oxygenase: protective gene or Trojan horse. Nat Med 4: 1364-1365, 1998." l. [( l5 S3 g1 B9 u6 U

& J; N( z. B; r
6 C4 c2 }" a% B
$ ^( B6 g" [: c  a/ bPolte T, Oberle S, and Schroder H. The nitric oxide donor SIN-1 protects endothelial cells from tumor necrosis factor- -mediated cytotoxicity: possible role for cyclic GMP and heme oxygenase. J Mol Cell Cardiol 29: 3305-3310, 1997.
3 B6 F9 S$ t) k! l( A7 @. V# {& H" R& l) `) [' T

" d5 c+ h  p: @! t
! Y* r7 j! W/ ?Poss KD and Tonegawa S. Reduced stress defense in heme oxygenase 1-deficient cells. Proc Natl Acad Sci USA 94: 10925-10930, 1997.- l1 f/ H) u- B* b

" {9 g- V" {7 W3 E; T& j8 x* @4 ~$ }5 [4 A. v
2 L. E* C: o# q8 a, b
Poss KD and Tonegawa S. Heme oxygenase 1 is required for mammalian iron reutilization. Proc Natl Acad Sci USA 94: 10919-10924, 1997.8 A, K; B/ X5 l* R' p; s6 O

: t0 u4 Z6 f5 j+ |# g
  a2 ^4 J" U7 X* Y! j+ d
# }6 r5 r& u& I9 x8 {2 o0 H( hRamesh G and Reeves WB. TNF- mediates chemokine and cytokine expression and renal injury in cisplatin nephrotoxicity. J Clin Invest 110: 835-842, 2002.7 Z% [  |. g+ m- g# D& R2 Y
8 g( S3 D: A. Y! x' l) Y

' }5 l- Y+ I% g' e$ V" H4 K( V" V
Rizzardini M, Terao M, Falciani F, and Cantoni L. Cytokine induction of haem oxygenase mRNA in mouse liver. Interleukin 1 transcriptionally activates the haem oxygenase gene. Biochem J 290: 343-347, 1993.8 _! X' t( i; v6 I# N4 y: q
; L# ~+ N" p, S4 N

3 F4 ~8 `6 a$ L  v* o. U
& N6 W8 s6 ?" O) H7 N; ]2 _' @% DRosenberger C, Griethe W, Gruber G, Wiesener M, Frei U, Bachmann S, and Eckardt KU. Cellular responses to hypoxia after renal segmental infarction. Kidney Int 64: 874-886, 2003.
: C" m, a0 f, Q5 |# M# e/ E( ^
9 o- C( @. z. f' r+ U7 n+ C
! v5 i0 c  B! d0 H  s& F* _. p& J: g  ~3 S
Ryter SW and Choi AM. Heme oxygenase-1: molecular mechanisms of gene expression in oxygen-related stress. Antioxid Redox Signal 4: 625-632, 2002.$ `& Z8 j% {% X6 U% |- P9 x

- {2 I# b2 r  Z- \& L
+ C4 I1 i; f( a+ |
- b9 U. a+ n+ B! ?/ P' g8 lSalinas M, Diaz R, Abraham N, Ruiz De Galarreta CM, and Cuadrado A. Nerve growth factor protects against 6-hydroxydopamine-induced oxidative stress by increasing expression of heme oxygenase-1 in a phosphatidylinositol 3 kinase-dependent manner. J Biol Chem 278: 13898-13904, 2003.# {/ [+ G' f3 [1 e) s3 \5 ~0 U% N

0 b. Q4 X% ], [4 j7 S' a
1 l8 J7 F! m6 p5 \
5 C5 X+ K# ?: O! W* g6 F: bSato K, Balla J, Otterbein L, Smith RN, Brouard S, Lin Y, Csizmadia E, Sevigny J, Robson SC, Vercellotti G, Choi AM, Bach FH, and Soares MP. Carbon monoxide generated by heme oxygenase-1 suppresses the rejection of mouse-to-rat cardiac transplants. J Immunol 166: 4185-4194, 2001.
+ j( q: O, @5 [6 M4 z- p+ V' A2 m8 ?1 k: i2 s% c1 b4 \( t
- ^) G+ ^( O2 Z* L8 G8 a

' s9 ?  ^. D: u1 ]Sato M, Ishizawa S, Yoshida T, and Shibahara S. Interaction of upstream stimulatory factor with the human heme oxygenase gene promoter. Eur J Biochem 188: 231-237, 1990.
* L% I) |+ l/ t% H# I$ B7 [' P4 {+ A8 _3 V; N- q. o

# n; S' d9 @0 n5 _  g* x0 f5 N' a9 E2 q2 P5 d+ [: ~: r9 l
Sharma VK, Bologa RM, Xu GP, Li B, Mouradian J, Wang J, Serur D, Rao V, and Suthanthiran M. Intragraft TGF- 1 mRNA: a correlate of interstitial fibrosis and chronic allograft nephropathy. Kidney Int 49: 1297-1303, 1996.
$ {. g5 r+ f; D4 G
0 O5 I7 c& ?. S: t3 _2 D; q3 a3 h  g5 H! p5 A/ g
8 G$ G. \' x3 o& R# u. X4 r
Shi Y, Wang YF, Jayaraman L, Yang H, Massague J, and Pavletich NP. Crystal structure of a Smad MH1 domain bound to DNA: insights on DNA binding in TGF- signaling. Cell 94: 585-594, 1998.
' m& W: N9 m6 O1 a4 W# {1 x$ n. g, C& l
" a9 }& G8 I5 _* ]0 E, j

1 `' G% q; H( l$ P. }% mShibahara S, Kitamuro T, and Takahashi K. Heme degradation and human disease: diversity is the soul of life. Antioxid Redox Signal 4: 593-602, 2002.6 U' r# Z+ N) n, n1 w7 [' r
1 d4 N% S: _( c6 \$ B, N
9 B+ F- j3 q- Y) @+ \' b

; L- q0 e3 {9 b/ m, C  C: rShibahara S, Muller RM, and Taguchi H. Transcriptional control of rat heme oxygenase by heat shock. J Biol Chem 262: 12889-12892, 1987.
- A) U- U( m, w! k" ^3 d
: ]: P; e' ~8 q5 c2 P- `: F% n1 V; R7 f
% w7 L6 F& l! o4 D: ^% P+ S( ^4 B
Shibahara S, Nakayama M, Kitamuro T, Udono-Fujimori R, and Takahashi K. Repression of heme oxygenase-1 expression as a defense strategy in humans. Exp Biol Med (Maywood) 228: 472-473, 2003.! q0 ?" y+ U2 K

' R/ E+ _9 L) R1 P. k
8 R4 _2 G/ a& Z8 X8 u
! m" Z! B. C' }, w5 a  o, \Shimizu H, Takahashi T, Suzuki T, Yamasaki A, Fujiwara T, Odaka Y, Hirakawa M, Fujita H, and Akagi R. Protective effect of heme oxygenase induction in ischemic acute renal failure. Crit Care Med 28: 809-817, 2000.* f$ K7 }9 w- o' I* K7 T" A

, j! {% l& }, [1 R0 |% a7 d! u; T# M  T& p3 j- U1 ~
' [5 J! D% v; i9 `1 }9 l
Shiraishi F, Curtis LM, Truong L, Poss K, Visner GA, Madsen K, Nick HS, and Agarwal A. Heme oxygenase-1 gene ablation or expression modulates cisplatin-induced renal tubular apoptosis. Am J Physiol Renal Physiol 278: F726-F736, 2000.
3 d. E4 X4 f% _$ ^- X1 g+ |5 F' z# Y# X& `
/ r& m, M- b' v* m

/ q7 o- P) q3 B- b' k! S% K/ oSnyder SH and Baranano DE. Heme oxygenase: a font of multiple messengers. Neuropsychopharmacology 25: 294-298, 2001.  d3 z' c7 ?) U# k0 c

5 y. G9 o% ]- ^  T! D8 s- S/ z6 N1 Y! C3 |- @

0 w8 `  `5 W: l9 p0 U/ |6 TSoares MP, Lin Y, Anrather J, Csizmadia E, Takigami K, Sato K, Grey ST, Colvin RB, Choi AM, Poss KD, and Bach FH. Expression of heme oxygenase-1 can determine cardiac xenograft survival. Nat Med 4: 1073-1077, 1998.
, N; a; G+ [* S5 E# t; I
0 h) Q$ V9 Z4 _" E- N; `$ O% T1 ?& a1 ^7 J6 K4 R
3 g  a, ]' y! x
Sporn MB and Roberts AB. Transforming growth factor- : recent progress and new challenges. J Cell Biol 119: 1017-1021, 1992.
5 B: j: U' k) R+ Y9 h
- n  J) c# L# E$ W& c9 R2 @  |1 l' v

: @$ E! @3 d! [9 \( oStewart D, Killeen E, Naquin R, Alam S, and Alam J. Degradation of transcription factor Nrf2 via the ubiquitin-proteasome pathway and stabilization by cadmium. J Biol Chem 278: 2396-2402, 2003.
; [4 l8 F/ }/ l8 M/ [9 b: u" y# P7 w+ U8 L; y5 y

; w6 c) e7 o7 |2 S& R8 E- U1 O' \, X0 g8 t! b4 l- {* D
Stocker R. Induction of haem oxygenase as a defence against oxidative stress. Free Radic Res Commun 9: 101-112, 1990.
* l: Y$ L' M3 x- N7 L* q+ _9 ^$ n( Q; G/ {" K- B

) U- W8 k. {: o3 b- s. G; d# E
6 x3 o9 n8 _1 q9 r! Q$ l" C1 OStocker R, Yamamoto Y, McDonagh AF, Glazer AN, and Ames BN. Bilirubin is an antioxidant of possible physiological importance. Science 235: 1043-1046, 1987.
8 F: K/ T7 E& |
6 _  D# S  m- n! [5 K/ V* {2 Z" Y) G7 c0 n. p6 a5 W, ~) Q9 R2 O1 T

9 W. B2 _1 c' CStroschein SL, Wang W, and Luo K. Cooperative binding of Smad proteins to two adjacent DNA elements in the plasminogen activator inhibitor-1 promoter mediates transforming growth factor -induced smad-dependent transcriptional activation. J Biol Chem 274: 9431-9441, 1999.! b# p/ J# h$ D
7 N2 Q  R. k8 n0 X) n% b" r2 i
% \0 D& R5 |1 L& l( v8 B
! K6 f0 t' k1 J! {- e1 e- ]
Sun J, Hoshino H, Takaku K, Nakajima O, Muto A, Suzuki H, Tashiro S, Takahashi S, Shibahara S, Alam J, Taketo MM, Yamamoto M, and Igarashi K. Hemoprotein Bach1 regulates enhancer availability of heme oxygenase-1 gene. EMBO J 21: 5216-5224, 2002.
, v0 Z: {/ h6 g  x5 I. G( t: X, A
0 p% w$ b$ H1 o9 t- l
% f" j( `( ?% Z: T  l
Suttner DM and Dennery PA. Reversal of HO-1 related cytoprotection with increased expression is due to reactive iron. FASEB J 13: 1800-1809, 1999.
: D2 C  G# m8 K
- B& ]2 j  F* a3 \8 Y- z* @8 g! @$ m3 p: E9 J

% [4 I9 @+ }* l! YTacchini L, Dansi P, Matteucci E, and Desiderio MA. Hepatocyte growth factor signalling stimulates hypoxia inducible factor-1 (HIF-1) activity in HepG2 hepatoma cells. Carcinogenesis 22: 1363-1371, 2001.
" g0 u* g& U9 a/ y& k
4 w, a2 Q- `! B* f/ Y9 r. Q  I- V. U* |* _( [

0 u, q* T9 s5 E7 N" r0 ^* O' iTakahashi K, Hara E, Ogawa K, Kimura D, Fujita H, and Shibahara S. Possible implications of the induction of human heme oxygenase-1 by nitric oxide donors. J Biochem (Tokyo) 121: 1162-1168, 1997.
1 h# w2 }7 P+ A" E9 c/ D$ _
: e: U' F) X1 q6 O+ }3 w: X0 ?0 B5 F" s. k- r
8 P% d0 T% A$ z1 P
Takahashi K, Nakayama M, Takeda K, Fujia H, and Shibahara S. Suppression of heme oxygenase-1 mRNA expression by interferon- in human glioblastoma cells. J Neurochem 72: 2356-2361, 1999.6 M5 J7 ~3 r) }( l
" B8 G! q: T0 D
; O2 F) y2 H( k6 C  P

1 K; M' q, S6 R2 v! Z( {# VTakahashi S, Matsuura N, Kurokawa T, Takahashi Y, and Miura T. Co-operation of the transcription factor hepatocyte nuclear factor-4 with Sp1 or Sp3 leads to transcriptional activation of the human haem oxygenase-1 gene promoter in a hepatoma cell line. Biochem J 367: 641-652, 2002.9 ~  K  o/ n- Q* O; u/ e) F1 U6 L* M

# K: z0 X0 Q* o# L' C' [+ u7 P5 y1 [) X) D, r1 t3 N2 |
5 p9 J7 c3 Q- t9 J
Takahashi S, Takahashi Y, Ito K, Nagano T, Shibahara S, and Miura T. Positive and negative regulation of the human heme oxygenase-1 gene expression in cultured cells. Biochim Biophys Acta 1447: 231-235, 1999.
5 |9 [3 s$ l3 V5 f% Q0 V8 A& h3 \- u; H, h
$ f6 F# p+ J6 ?( c: y! j. ^
1 O: {, {$ ^8 X
Takahashi S, Takahashi Y, Yoshimi T, and Miura T. Oxygen tension regulates heme oxygenase-1 gene expression in mammalian cell lines. Cell Biochem Funct 16: 183-193, 1998. <a href="/cgi/external_ref?access_num=10.1002/(SICI)1099-0844(199809)16:3. d/ [% F$ `. T6 z

* {! U7 K) ?% y6 Z. Q! r3 d) q+ U
, F" U; g3 g  Z2 B5 U$ z' d
3 ~7 C1 G( H, u0 d- v/ r/ w7 L; }Takeda K, Fujita H, and Shibahara S. Differential control of the metal-mediated activation of the human heme oxygenase-1 and metallothionein IIA genes. Biochem Biophys Res Commun 207: 160-167, 1995.( ~1 l6 r) E; W
; e5 e& _  A- b! a
% U5 C$ Q: D6 E8 ]4 r

) W# ]* |! A1 d9 b3 H* E& rTakeda K, Ishizawa S, Sato M, Yoshida T, and Shibahara S. Identification of a cis-acting element that is responsible for cadmium-mediated induction of the human heme oxygenase gene. J Biol Chem 269: 22858-22867, 1994.
  A$ x# b, \( `( j* g$ D1 K$ N: c; `; {7 l  L

. H- v7 a, }1 K$ b/ v$ I
0 H" X5 f7 j+ t" p  k3 J3 zTenhunen R, Marver HS, and Schmid R. The enzymatic conversion of heme to bilirubin by microsomal heme oxygenase. Proc Natl Acad Sci USA 61: 748-755, 1968." n$ L/ L* v) _+ n. v( ~% r3 J
5 ^4 v: D4 U" Y" B4 r( T7 q

7 d; y9 b0 R* A( V4 h$ a
2 v, {5 R1 }4 B  YTenhunen R, Marver HS, and Schmid R. Microsomal heme oxygenase. Characterization of the enzyme. J Biol Chem 244: 6388-6394, 1969.
4 n- G9 i3 u: c, z7 v' X
( R3 {, s5 l9 s! @# t7 |' j/ I2 b/ u* {& ^/ a0 R9 j
- @! T& I* H! h' |2 g2 \5 C) v
Terry CM, Clikeman JA, Hoidal JR, and Callahan KS. Effect of tumor necrosis factor- and interleukin-1 on heme oxygenase-1 expression in human endothelial cells. Am J Physiol Heart Circ Physiol 274: H883-H891, 1998.
9 W7 t$ c8 G: Z% G3 ?6 M% W' d/ m
3 v- E2 x3 u* Y$ G  d$ t8 w6 b8 h# |! S

; r* m/ ?: ]4 V% Z. Q# y: ITerry CM, Clikeman JA, Hoidal JR, and Callahan KS. TNF- and IL-1 induce heme oxygenase-1 via protein kinase C, Ca 2 , and phospholipase A 2 in endothelial cells. Am J Physiol Heart Circ Physiol 276: H1493-H1501, 1999.( k0 X- F, l) J0 P3 }. C# ?8 N
$ T0 c0 E/ R! u: w
" g+ }# h$ I; v9 P0 l9 I

: X/ z+ D) D# s5 {$ z/ jTetsuka T, Daphna-Iken D, Srivastava SK, and Morrison AR. Regulation of heme oxygenase mRNA in mesangial cells: prostaglandin E 2 negatively modulates interleukin-1-induced heme oxygenase-1 mRNA. Biochem Biophys Res Commun 212: 617-623, 1995.
9 Q* Q6 J' e$ f& M/ U. z# E4 `3 W/ T* q% r0 U& t0 ~" y  N' I9 R

- x: x& k% g8 v$ M# a  W
1 q; b* _- k" f. f5 d7 K& [Thiemermann C. Inhaled CO: deadly gas or novel therapeutic? Nat Med 7: 534-535, 2001.
- K! c9 T2 Q5 L2 e( \7 |" E
1 S9 l2 |# t, A
3 C( M( C- \+ D* u% I! }, v2 |. d; H$ `9 z" X
Thom SR, Fisher D, Xu YA, Notarfrancesco K, and Ischiropoulos H. Adaptive responses and apoptosis in endothelial cells exposed to carbon monoxide. Proc Natl Acad Sci USA 97: 1305-1310, 2000.6 H( m% Y: [& j0 L
8 e/ s* s4 B6 y4 Q" u
! F8 F$ y/ d# F
: P' N: @2 w* g8 w
Tian W, Bonkovsky HL, Shibahara S, and Cohen DM. Urea and hypertonicity increase expression of heme oxygenase-1 in murine renal medullary cells. Am J Physiol Renal Physiol 281: F983-F991, 2001.
( C3 T9 D, p  N# z, x& K% Y$ w1 Y1 t/ O' A/ P7 a' K

$ k0 q: x0 ^+ O7 D6 p. m1 Z& U! Y$ g0 X1 R
Tyrrell R. Redox regulation and oxidant activation of heme oxygenase-1. Free Radic Res 31: 335-340, 1999.  R- }. ^! ]7 v- B3 ?
+ F) }& S8 q" ?! c2 z
' o; i; ], r, f- l
2 ~3 \, C+ U' X$ ~- P3 y
Tyrrell RM, Applegate LA, and Tromvoukis Y. The proximal promoter region of the human heme oxygenase gene contains elements involved in stimulation of transcriptional activity by a variety of agents including oxidants. Carcinogenesis 14: 761-765, 1993.
, t7 z; ]" Z, M; W
! C8 s6 y3 ]9 U! s9 i; C& I; n
) q  Y- A5 T' [& A
+ [% v( ]/ D3 K& |8 |$ NVogt BA, Shanley TP, Croatt A, Alam J, Johnson KJ, and Nath KA. Glomerular inflammation induces resistance to tubular injury in the rat. A novel form of acquired, heme oxygenase-dependent resistance to renal injury. J Clin Invest 98: 2139-2145, 1996.
, M% b; M8 _, z) {. q5 I8 ^' B4 D3 y; \" T- H4 Z$ Z8 F. c) L5 M/ q

: F3 N/ U( `) L2 |  u# u" ^' g0 \& U( N
Von Gersdorff G, Susztak K, Rezvani F, Bitzer M, Liang D, and Bottinger EP. Smad3 and Smad4 mediate transcriptional activation of the human Smad7 promoter by transforming growth factor. J Biol Chem 275: 11320-11326, 2000.
9 o* F  k/ K, }& s6 P1 G6 K5 b
- t/ C2 U( ?5 [* c8 u) O
) F* G, v$ B# Z, N* |2 I: `. U. }4 C& T1 S3 C" z3 F" ]8 n# Z" F
Wagener FA, da Silva JL, Farley T, de Witte T, Kappas A, and Abraham NG. Differential effects of heme oxygenase isoforms on heme mediation of endothelial intracellular adhesion molecule 1 expression. J Pharmacol Exp Ther 291: 416-423, 1999.1 X" W( p) x4 d  A

+ r& v; z  D9 n; p" {9 n, z
' I4 u0 B, ?, W! ^" u9 S2 x- p$ G
0 {" Z/ P, s' y; s2 p4 b8 A" d9 LWagner CT, Durante W, Christodoulides N, Hellums JD, and Schafer AI. Hemodynamic forces induce the expression of heme oxygenase in cultured vascular smooth muscle cells. J Clin Invest 100: 589-596, 1997.
3 k; X2 }! Q4 Q+ O% x3 B1 ^  o) }. k
$ o4 d# v( e. G- f5 f2 c$ Q8 X6 A0 p  A" r/ s* q. w# i
5 r9 C4 \/ ?0 d- f: c# A  E' s
Wang LJ, Lee TS, Lee FY, Pai RC, and Chau LY. Expression of heme oxygenase-1 in atherosclerotic lesions. Am J Pathol 152: 711-720, 1998.% Y/ h0 o; |6 L
# |- X% f& Q: w

$ V3 z  p7 U9 k% h
( {& N* t- T7 a$ z, CWiesel P, Patel AP, Carvajal IM, Wang ZY, Pellacani A, Maemura K, DiFonzo N, Rennke HG, Layne MD, Yet SF, Lee ME, and Perrella MA. Exacerbation of chronic renovascular hypertension and acute renal failure in heme oxygenase-1-deficient mice. Circ Res 88: 1088-1094, 2001.
  a) K; i6 \. T, R( j: Q0 n: s1 }+ }. \
1 {3 R1 _- ]/ S4 x* T: B

" P: |& D3 g, j% m8 d" U: Y$ tWild AC, Moinova HR, and Mulcahy RT. Regulation of -glutamylcysteine synthetase subunit gene expression by the transcription factor Nrf2. J Biol Chem 274: 33627-33636, 1999.
6 t4 k! y/ C; X, a, e3 g2 p; w5 U
5 R, w( w# W/ E. g* k* R/ X5 T. c
8 [  \; G: G$ L: G" ^- O0 K* R
% M+ [* z; u( w$ f+ {Wood SM, Wiesener MS, Yeates KM, Okada N, Pugh CW, Maxwell PH, and Ratcliffe PJ. Selection and analysis of a mutant cell line defective in the hypoxia-inducible factor-1 -subunit (HIF-1 ). Characterization of HIF-1 -dependent and -independent hypoxia-inducible gene expression. J Biol Chem 273: 8360-8368, 1998.
8 A, k$ U$ \: t5 S; `5 ^6 v5 _$ e: C# h( t9 {+ W; {

; g- o, G" I) k2 o/ I" V
/ \6 f# W% a+ FYachie A, Niida Y, Wada T, Igarashi N, Kaneda H, Toma T, Ohta K, Kasahara Y, and Koizumi S. Oxidative stress causes enhanced endothelial cell injury in human heme oxygenase-1 deficiency. J Clin Invest 103: 129-135, 1999.7 t3 h7 C4 K* [  Q" ]( Y+ C7 r

  O' K; N2 H$ n( ]8 r, o* y" M1 y( I) `; _! n

9 T! E% V+ X6 ^, o  r' g! ]Yamada N, Yamaya M, Okinaga S, Nakayama K, Sekizawa K, Shibahara S, and Sasaki H. Microsatellite polymorphism in the heme oxygenase-1 gene promoter is associated with susceptibility to emphysema. Am J Hum Genet 66: 187-195, 2000.
8 z( ?2 a) c  q6 ~- r2 f
! Z4 _5 i/ w4 w
5 v! G* n/ S1 P' F2 a
6 h7 {, |, J0 |) eYamamoto T, Noble NA, Cohen AH, Nast CC, Hishida A, Gold LI, and Border WA. Expression of transforming growth factor- isoforms in human glomerular diseases. Kidney Int 49: 461-469, 1996.
& i3 z1 D- a2 C  v
! d- h) F) m, [
' E& ^7 U4 Z7 `9 o/ }5 T
" @5 U  x; K. e+ i* p( ~. OYang ZZ and Zou AP. Transcriptional regulation of heme oxygenases by HIF-1 in renal medullary interstitial cells. Am J Physiol Renal Physiol 281: F900-F908, 2001.  ]" ]9 g  @2 ^: |0 H" v" d

1 D9 ~  ]# g5 F5 r" E( h8 h7 F. [# g
& Y* G/ w5 D8 k7 @; p, Z* V! z5 Z/ }( g' U# Y
Yet SF, Pellacani A, Patterson C, Tan L, Folta SC, Foster L, Lee WS, Hsieh CM, and Perrella MA. Induction of heme oxygenase-1 expression in vascular smooth muscle cells. A link to endotoxic shock. J Biol Chem 272: 4295-4301, 1997.
4 Z$ V# {8 j  K) K, R& Y+ P
! {0 E1 g1 i7 b! ?5 R! O/ R
6 @: r: N: R7 L5 {) k7 N6 m7 ]) C+ H: L
Yet SF, Perrella MA, Layne MD, Hsieh CM, Maemura K, Kobzik L, Wiesel P, Christou H, Kourembanas S, and Lee ME. Hypoxia induces severe right ventricular dilatation and infarction in heme oxygenase-1 null mice. J Clin Invest 103: R23-R29, 1999.
; {2 h% H: \3 ~* w  S! n
: `# q6 M8 l. z# q9 I  h$ s9 |4 u
# U9 |* n5 k9 t& q- o
Yoneya R, Ozasa H, Nagashima Y, Koike Y, Teraoka H, Hagiwara K, and Horikawa S. Hemin pretreatment ameliorates aspects of the nephropathy induced by mercuric chloride in the rat. Toxicol Lett 116: 223-229, 2000.' Z- `4 F4 c  f' ~

* g: e4 \! z# B* y, n) l# U: i0 G2 K" [2 @6 h8 e4 \
# c8 H, _; `  c$ n7 P% c
Yoshida T, Biro P, Cohen T, Muller RM, and Shibahara S. Human heme oxygenase cDNA and induction of its mRNA by hemin. Eur J Biochem 171: 457-461, 1988.
3 b$ S. {; ]2 R
' E2 Q+ {/ u) a$ B- c# C- q
% D- X8 j( v% m* Y, [
7 K8 S5 h  c- [( O" _3 _5 ?/ YYu L. Nitric oxide in acute renal failure: foe or friend? Kidney Int Suppl 61: S39-S40, 1997.
* j- G$ r4 y; q8 W1 k& {; x" C
% X) @7 a8 y- k5 S4 C( w% o
+ {* W% r9 |- G
9 J, j5 n, D) g+ l6 _2 d& aYu L, Gengaro PE, Niederberger M, Burke TJ, and Schrier RW. Nitric oxide: a mediator in rat tubular hypoxia/reoxygenation injury. Proc Natl Acad Sci USA 91: 1691-1695, 1994.
$ e) V/ R# P. r; `. j7 E7 n) ~; m* `' R

+ }2 Q& S9 ]1 b, r: \& m( `: I; |; w+ {  M% d. b
Yu L, Hebert MC, and Zhang YE. TGF- receptor-activated p38 MAP kinase mediates Smad-independent TGF- responses. Embo J 21: 3749-3759, 2002.
1 g' m% ~0 a  p! D' O$ L4 g- P4 V  q/ m% }

+ T7 {$ W* E9 n! x/ _5 K" L2 q2 k5 G" O7 J
Yu R, Chen C, Mo YY, Hebbar V, Owuor ED, Tan TH, and Kong AN. Activation of mitogen-activated protein kinase pathways induces antioxidant response element-mediated gene expression via a Nrf2-dependent mechanism. J Biol Chem 275: 39907-39913, 2000./ v; ^; t) N8 ]7 V5 q
5 F9 ?4 P: J, ~9 V. A) g  J( H

. ?% ^. Z1 R  N) S! T- ]1 k* V$ X( U5 u' \
Zager RA. Heme protein-induced tubular cytoresistance: expression at the plasma membrane level. Kidney Int 47: 1336-1345, 1995.5 J1 j0 i9 B1 U) s6 ^

* x* n& O1 v4 K
: d) l0 S6 B3 \% N+ h, n; h# e" d6 t- f$ B3 D: |3 l( {
Zager RA, Burkhart KM, Conrad DS, and Gmur DJ. Iron, heme oxygenase, and glutathione: effects on myohemoglobinuric proximal tubular injury. Kidney Int 48: 1624-1634, 1995.
% l! p$ I  j4 D, C: e% |7 O. s, Q- r5 ?. \
: D7 [! K: K) E# |6 N1 b
) ~) K, j3 G. f( ~# @$ g8 D9 j
Zawel L, Dai JL, Buckhaults P, Zhou S, Kinzler KW, Vogelstein B, and Kern SE. Human Smad3 and Smad4 are sequence-specific transcription activators. Mol Cell 1: 611-617, 1998.
: N1 x  b; b$ I9 {) G' X% j. B+ L
' K* u" o$ ]8 j: U# ^
6 C- p8 n( ?% F/ H
" ~, c( a: s! c8 nZhang J and Piantadosi CA. Mitochondrial oxidative stress after carbon monoxide hypoxia in the rat brain. J Clin Invest 90: 1193-1199, 1992.- _9 A. n$ a/ i
- K! l0 O. u0 R. J" ?
) X, T* b- b; L+ n( ?
4 m% _  ~1 v# q& m# o- ^+ Q& x
Zhang W, Contag PR, Hardy J, Zhao H, Vreman HJ, Hajdena-Dawson M, Wong RJ, Stevenson DK, and Contag CH. Selection of potential therapeutics based on in vivo spatiotemporal transcription patterns of heme oxygenase-1. J Mol Med 80: 655-664, 2002.
% k8 G1 N1 {" j) L; ~6 b; B$ u- L
1 E: f8 ^+ |3 [- Q7 y# K4 w( r! @! U; |$ {3 D! h: T3 m% C7 s

4 a9 X1 r& P& Z# \- vZhu HJ, Iaria J, and Sizeland AM. Smad7 differentially regulates transforming growth factor -mediated signaling pathways. J Biol Chem 274: 32258-32264, 1999.
, \- x5 u0 U! d- d8 ~$ ~
8 [. K* H& ]8 I! D8 q4 ?6 ]" I# p* d9 p8 ~$ [/ @. ?

, u& V8 X& h' ~3 X6 n% f# D. q( f7 J7 dZou AP, Billington H, Su N, and Cowley AW Jr. Expression and actions of heme oxygenase in the renal medulla of rats. Hypertension 35: 342-347, 2000.& {# w/ B4 @; \; s# O# K
+ o6 }0 W% i2 `! y, {1 N  _
) |* G$ r! {& W( R
6 r  ?' Z6 ]* c1 y6 F& a# c
Zuckerbraun BS and Billiar TR. Heme oxygenase-1: a cellular Hercules. Hepatology 37: 742-744, 2003.

Rank: 2

积分
132 
威望
132  
包包
1727  
沙发
发表于 2015-5-28 11:10 |只看该作者
每天早上起床都要看一遍“福布斯”富翁排行榜,如果上面没有我的名字,我就去上班……  

Rank: 2

积分
88 
威望
88  
包包
1897  
藤椅
发表于 2015-6-14 14:27 |只看该作者
貌似我真的很笨????哎  

Rank: 2

积分
80 
威望
80  
包包
1719  
板凳
发表于 2015-7-13 20:18 |只看该作者
干细胞之家微信公众号
我毫不犹豫地把楼主的这个帖子收藏了  

Rank: 2

积分
104 
威望
104  
包包
1772  
报纸
发表于 2015-8-31 08:54 |只看该作者
晕死也不多加点分  

Rank: 2

积分
104 
威望
104  
包包
1772  
地板
发表于 2015-9-1 11:01 |只看该作者
世界上那些最容易的事情中,拖延时间最不费力。  

Rank: 2

积分
162 
威望
162  
包包
1724  
7
发表于 2015-9-20 04:31 |只看该作者
加油啊!!!!顶哦!!!!!支持楼主,支持你~  

Rank: 2

积分
129 
威望
129  
包包
1788  
8
发表于 2015-9-25 13:43 |只看该作者
小生对楼主之仰慕如滔滔江水连绵不绝,海枯石烂,天崩地裂,永不变心.  

Rank: 2

积分
61 
威望
61  
包包
1757  
9
发表于 2015-10-7 14:33 |只看该作者
干细胞研究非常有前途

Rank: 2

积分
166 
威望
166  
包包
1997  
10
发表于 2015-11-12 08:17 |只看该作者
不错啊! 一个字牛啊!  
‹ 上一主题|下一主题
你需要登录后才可以回帖 登录 | 注册
验证问答 换一个

Archiver|干细胞之家 ( 吉ICP备2021004615号-3 )

GMT+8, 2024-6-7 10:33

Powered by Discuz! X1.5

© 2001-2010 Comsenz Inc.